blood-brain barrier leakage after transient cerebral ischemia

Transkript

blood-brain barrier leakage after transient cerebral ischemia
Department of Neurology
Helsinki University Central Hospital
University of Helsinki
Helsinki, Finland
BLOOD-BRAIN BARRIER LEAKAGE
AFTER TRANSIENT CEREBRAL ISCHEMIA
Aysan Durukan Tolvanen
ACADEMIC DISSERTATION
To be publicly discussed with the permission of the Medical Faculty of the University
of Helsinki in Lecture Hall 3, Biomedicum Helsinki 1, Haartmaninkatu 8,
on January 31, 2014, at 12 noon.
Helsinki, 2014
SUPERVISORS
Docent Turgut Tatlisumak
Department of Neurology
Helsinki University Central Hospital
and Experimental MRI Laboratory, Biomedicum Helsinki
Docent Daniel Strbian
Department of Neurology
Helsinki University Central Hospital
and Experimental MRI Laboratory, Biomedicum Helsinki
REVIEWERS
Professor Olli Gröhn
Department of Neurobiology
A. I. Virtanen Institute for Molecular Sciences
University of Eastern Finland
Docent Jari Honkaniemi
Department of Neurology
University of Tampere
Tampere, Finland
OPPONENT
Professor Schäebitz Wolf-Rudiger
Department of Neurology
University of Mϋnster
Mϋnster, Germany
Aysan Durukan Tolvanen, M.D.
[email protected]
Cover photo: Brain party by Geneviève Gauckler,
reproduced with the kind permission of Geneviève Gauckler.
ISBN 978-952-10-9698-3 (paperback)
ISBN 978-952-10-9699-0 (PDF)
http://ethesis.helsinki.fi
Helsinki University Print
Helsinki, Finland 2014
2
“Imagination is more important than knowledge.”
Albert Einstein
3
CONTENTS
LIST OF ORIGINAL PUBLICATIONS and AUTHOR’S CONTRIBUTION.................6
ABBREVIATIONS.......................................................................................................7
ABSTRACT.................................................................................................................8
1 REVIEW OF THE LITERATURE............................................................................10
1.1
Ischemic stroke
........................................................................................10
1.1.1
Pathophysiology of ischemic stroke..................................................................11
1.1.1.1
Core and penumbra.....................................................................11
1.1.1.2
Ischemic cascade........................................................................13
1.1.2
Acute ischemic stroke therapy..........................................................................16
1.1.2.1
Recanalization.............................................................................16
1.1.2.2
Neuroprotection...........................................................................18
1.1.3
Magnetic resonance imaging (MRI) in acute ischemic stroke………………....19
1.1.3.1
Lesion evaluation by MRI............................................................20
1.1.3.2
BBB disruption: contrast-enhanced imaging...............................21
1.2
Experimental ischemic stroke………............................................................23
1.2.1
Why animal modeling.......................................................................................23
1.2.2
Major rodent models of ischemic stroke...........................................................23
1.2.2.1
Thromboembolic models.............................................................24
1.2.2.2
Suture occlusion of the MCA………............................................25
1.2.2.3
Other models...............................................................................26
1.2.3
Preconditioning.................................................................................................28
1.2.3.1
Ischemic tolerance......................................................................28
1.2.3.2
Hypoxic preconditioning..............................................................28
1.2.4
Outcome measures..........................................................................................29
1.2.4.1
Infarct volume..............................................................................30
1.2.4.2
Neurological status......................................................................30
1.2.5
Sources of variability in experimental ischemic stroke....................................31
1.3
Blood-brain barrier……................................................................................33
1.3.1 Structure and functions, neurovascular unit.....................................................33
1.3.1.1
Endothelial cells and pericytes....................................................34
1.3.1.2
Basal lamina................................................................................35
1.3.1.3
Tight junctions.............................................................................35
1.3.1.4
Adherens junctions......................................................................37
1.3.1.5
Astrocytes....................................................................................37
1.3.2
Methods to evaluate BBB permeability............................................................38
1.3.2.1
Qualitative methods.....................................................................38
1.3.2.1.1
Visualization of dye extravasation...............................................38
1.3.2.1.2
Visualization of contrast agent extravasation ………..................38
1.3.2.2
Quantitative methods..................................................................39
1.3.2.2.1
Colorimetric and fluorometric methods........................................39
1.3.2.2.2
Autoradiographic method............................................................39
1.3.2.2.3
Fluorescence methods................................................................40
1.3.2.2.4
Other methods.............................................................................40
1.3.2.2.5
Dynamic contrast-enhanced MRI (DCE-MRI).............................40
4
1.3.3
BBB disruption in experimental stroke..............................................................41
1.3.3.1
Theories of biphasic BBB disruption...........................................42
1.3.3.1
Continuous BBB disruption.........................................................43
1.4
Brain ischemia and stanniocalcin…….........................................................44
2 AIMS OF THE STUDY...........................................................................................45
3 MATERIALS AND METHODS...............................................................................46
3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
Animals........................................................................................................46
Anesthesia...................................................................................................46
Monitoring of physiological parameters.......................................................46
Study designs..............................................................................................47
Focal cerebral ischemia model....................................................................48
Hypoxic preconditioning...............................................................................49
Laser-Doppler flowmetry..............................................................................49
MRI studies..................................................................................................50
3.8.1
3.8.2
Patlak plotting.................................................................................................51
Imaging protocol.............................................................................................53
3.9 Neurological evaluation...............................................................................54
3.10 Tissue handling............................................................................................54
3.11 Ischemic lesion assessment.......................................................................54
3.11.1
3.11.2
3.12
MRI-based infarction..................................................................................... 54
TTC-based infarction......................................................................................55
Blood-brain barrier permeability assessments............................................55
3.12.1
Evans blue extravasation...............................................................................55
3.12.2
Contrast-enhanced MRI.................................................................................56
3.12.2.1
Percentage of enhancement of the ischemic lesion....................56
3.12.2.2
Contrast-to-noise ratio of the enhancement area........................56
3.12.2.3
Signal intensity change due to enhancement............................ 56
3.12.2.4
The blood-to-brain transfer constant of Gd-DTPA.......................56
3.13
3.14
Quantitative analyses of Stc1, Stc2, and Il-6 mRNA..................................57
Statistical analyses....................................................................................57
4 RESULTS...............................................................................................................58
5 DISCUSSION.........................................................................................................64
6 SUMMARY AND CONCLUSIONS.........................................................................71
ACKNOWLEDGMENTS............................................................................................72
REFERENCES..........................................................................................................74
ORIGINAL PUBLICATIONS.....................................................................................92
5
LIST OF ORIGINAL PUBLICATIONS and AUTHOR’S CONTRIBUTION
This thesis is based on the following publications, referred to in the text by their Roman
numerals:
I
Strbian D*, Durukan A*, Pitkonen M, Marinkovic I, Tatlisumak E, Pedrono E,
Abo-Ramadan U, Tatlisumak T. The blood-brain barrier is continuously open for several
weeks following transient focal cerebral ischemia. Neuroscience 2008; 153:175-181
II
Abo-Ramadan U*, Durukan A*, Pitkonen M, Marinkovic I, Pedrono E, Soinne L,
Strbian D, Tatlisumak T. Post-ischemic leakiness of the blood-brain barrier: a quantitative
and systematic assessment by Patlak plots. Exp Neurol 2009; 219:328-333
III
Durukan A*, Marinkovic I*, Pitkonen M, Abo-Ramadan U, Pedrono E, Soinne L,
Strbian D, Tatlisumak T. Post-ischemic blood-brain barrier leakage in rats: one-week followup by MRI. Brain Res 2009; 1280:158-165
IV
Durukan Tolvanen A, Westberg JA; Serlachius M, Chang AC-M ; Reddel RR,
Andersson LC; Tatlisumak T. Stanniocalcin 1 is important for poststroke functionality, but
dispensable for ischemic tolerance. Neuroscience 2013; 229: 49-54
*Equal contribution.
In Study I, Aysan Durukan Tolvanen performed most of the experiments, contributed to data
analysis and interpretation, and provided intellectual content to the manuscript. In Study II,
the author performed most of the experiments, contributed to data analysis and
interpretation, and wrote the manuscript. In Study III, Aysan Durukan Tolvanen performed
most of the experiments, contributed to data analysis and interpretation, and wrote the
manuscript. In Study IV, the author performed all experiments, except mRNA analyses,
analyzed and interpreted the data, and wrote the manuscript.
6
ABBREVIATIONS
ADC
Apparent diffusion coefficient
AIS
AQP4
BBB
BBBP
CBF
CT
DCE-MRI
DWI
EB
ET-1
FLAIR
Gd-DTPA
HIF-1
HPC
IL
IR-FLASH
JAM
Ki
MCA
MCAO
MMP
MRI
PWI
ROI
STC
STC1-/TJ
TTC
t-PA
T1-WI
WT
ZO
Acute ischemic stroke
Aquaporine-4
Blood-brain barrier
BBB permeability
Cerebral blood flow
Computed tomography
Dynamic contrast-enhanced MRI
Diffusion-weighted imaging
Evans blue
Endothelin-1
Fluid-attenuated inversion-recovery
Gadolinium diethylenetriaminepentaacetic acid
Hypoxia-inducible factor-1
Hypoxic preconditioning
interleukin
inversion recovery snapshot-fast low-angled shot
Junctional adhesion molecule
Blood-to-brain transfer rate constant of the contrast agent
Middle cerebral artery
MCA occlusion
Matrix metalloproteinase
Magnetic resonance imaging
Perfusion-weighted imaging
Region of interest
Stanniocalcin
STC1 knockout
Tight junctions
2,3,5-Triphenyltetrazolium chloride
Tissue-plasminogen activator
T1-weighted image
Wild type
Zonula occludens
7
ABSTRACT
Acute ischemic stroke (AIS) is a devastating disease leaving more than half of its victims
disabled and causing nearly 5% of all deaths worldwide. In large ischemic strokes, a major
cause of death is brain edema, which follows blood-brain barrier (BBB) leakage. The BBB
ensures brain homeostasis in health and disease by limiting the entry of harmful blood-borne
substances into the brain parenchyma. With a leaky BBB, the brain becomes devoid of
protection from detrimental components of the circulating blood.
The BBB leakage in animal models of ischemia–reperfusion has long been considered to be
biphasic; however, a considerable amount of discrepancies exist among the studies.
Knowing exact temporal changes of the BBB permeability (BBBP) is important for the
management of stroke patients. When the BBB is open, BBBP alleviating therapies would be
effective, neuroprotective or neurorestorative drugs would be introduced, and if the BBB is
closed these drugs would not enter the brain. Practical and reliable biomarkers of BBBP
status are needed.
Stanniocalcins (STCs) are widely expressed in the brain and STC-1 expression is elevated
in pathologies, such as hypoxia and focal ischemia. Recent data suggest a neuroprotective
role for STC-1 especially trough hypoxic preconditioning (HPC). No previous data
associate STC-1 and the BBB.
We systematically evaluated disruption of the BBB following ischemia-reperfusion in a rat
model of transient focal ischemia via suture occlusion of the middle cerebral artery for 90
min. Firstly (I, II), animals were allocated to 15 groups after reperfusion (25 min to 5 weeks).
Secondly (III), a group of animals were evaluated repeatedly from 2 h to 1 week after
reperfusion. BBBP to both small (gadolinium) (I, II, III), and large (Evans blue) (I)
molecules were quantified by magnetic resonance imaging and fluorescence, respectively.
Lastly, the contribution of STC-1 to HPC and the BBB was explored using STC-1 deficient
mice (STC-/-).
8
(I, II, III) After transient ischemia, the BBB leakage was continuous. Leakage to Evans Blue
persisted up to 3 weeks and to gadolinium up to 5 weeks. Evans blue leakage slightly
decreased at 36 and 72 h, gadolinium leakage was lesser at 25 min, 3 and 4 weeks. (IV) In
STC-/- mice, HPC was effective in reducing lesion size, but these mice scored worse than
wild type littermates. BBBP to Evans blue was not increased in STC-/- mice; neither under
normal conditions, nor after hypoxia.
To conclude, transient focal ischemia in rats triggers a continuous BBB leakage lasting for
several weeks. Until the final closure of the BBB, no earlier transient closure occurs. This
finding indicates a long therapeutic window opportunity in respect to BBB passage of drugs
to treat stroke. BBBP imaging method used in these studies may be easily translated to
clinics. STC-1 is not obligatory for hypoxic preconditioning and is not a determining
component of the BBB. Yet, STC-1 is important for preservation of neurological function after
transient ischemia.
9
1 REVIEW OF THE LITERATURE
1.1
ISCHEMIC STROKE
Stroke is a devastating disease. At the year 2010, stroke ranked the second leading cause of
death, responsible for 8.9% of all deaths worldwide.1 It is estimated that each year nearly 6
million people die from stroke and another 5 million are left dependant on others. 2 Although
the incidence of stroke is declining in many industrial countries due to improved preventive
treatment, absolute number of strokes increases because of ageing. The world's 65-andolder population is estimated to triple by midcentury, from 516 million in 2009 to 1.53 billion in
2050.3 Europeans will likely continue to be the oldest people in the world: by 2050, 29
percent of Europe’s population is estimated to be 65 years and older.3 Stroke burden
calculated as disability-adjusted life years (a combination of years of life lost due to death
and years of disability) is projected to rise by nearly two-fold from year 1990 to year 2020. In
2001, disability-adjusted life years due to stroke were around 72 millions.4 In Finland stroke
absorbs 7% of the health care budget.5 In the United States, the total cost of stroke in 2008
was 34.3 billion and doubled in 2010, and the mean lifetime cost of ischemic stroke is
estimated at $140,048 per patient.6
Ischemic stroke accounts for 80 to 85% of all cases of stroke and results from a thrombotic
or embolic occlusion of a cerebral artery. Over half of all ischemic strokes occur in the middle
cerebral artery (MCA) territory.7 Etiologically atherosclerosis, embolism of cardiac origin, and
small artery disease explain majority of cases.8 A minority is caused by dissections and over
100 other causes. Eighty per cent of all ischemic strokes are preventable. Key risk factors for
ischemic stroke are arterial hypertension, smoking, and hypercholesterolemia. Other risk
factors can be classified as: traits (advanced age, male gender, African-American race,
family history), medical conditions (diabetes mellitus, ischemic heart disease, atrial
fibrillation, carotid artery disease, heart failure, peripheral arterial disease, thrombotic
disorders, chronic kidney disease, sleep apnea), life style choices (obesity, excessive alcohol
use, insufficient sleep, physical inactivity), and use of certain drugs (oral contraceptives,
hormone replacement therapy, illicit drugs).
10
Intravenous thrombolysis with tissue plasminogen activator (t-PA) opened a new era in the
management of acute ischemic stroke (AIS). Besides thrombolysis, the effect of stroke unit
care, aspirin, and hemicraniectomy are evidence-based. Secondary prevention requires
control of modifiable risk factors and etiology-oriented treatment (such as anticoagulation in
atrial fibrillation, thrombophilia, and dissection, or carotid endarterectomy in significant largeartery atherosclerosis).
1.1.1
Pathophysiology of ischemic stroke
The brain, although a tiny organ in size (2% of body weight), uses nearly one fifth of body’s
oxygen and blood supply. Stores of energy lack in the brain making it highly dependent on
continuous cerebral blood flow (CBF). Therefore, within minutes of cessation of blood supply
to a territory of the brain, a complex sequence of pathophysiological spatial and temporal
events (ischemic cascade) occurs.
1.1.1.1 Core and penumbra
Conventionally, core represents the irreversibly injured part of the ischemic lesion that
destined to infarction and penumbra represents the region that is dysfunctional but
salvageable if regional CBF is restored.9-12 Positron emission tomography techniques
revealed: 1) the core with a CBF of <12mL/100g per min and 2) the penumbra with a CBF
of 12 to 22mL/100g per min.13 A third benign oligemic area with CBF >22mL/100g per min
also appeared13 (Figure 1). This oligemic tissue probably maintains its function for a very
long time and is unlikely proceed to infarction.14 Animal studies using autoradiography to
measure local CBF defined the core as a ischemic zone where CBF reduced to 0% to 20%
of control, and the penumbra showed intermediate reductions of the flow, 20% to 40% of the
control.15 In the core, protein synthesis seizes related to ATP loss and irreversible translation
blockade. In the penumbra ATP is preserved, protein synthesis is decreased, heat shock
proteins are produced, and probably an unfolded protein response occurs.16
11
Figure 1 Schematic presentation of ischemic core and penumbra.
Pathologic outcome of focal ischemia is determined by two main factors: the degree of
ischemia (rate of CBF decrease) and the duration of ischemia (time from CBF decrease to its
recovery).17, 18 The probability of infarction is greater than 95% if early CBF falls below 25%
of control.19 Early reperfusion preserves penumbra, but if the blood flow is not restored, the
core extents to the entire penumbra. In most animal models, transient focal ischemia lasting
4 hours or more induces similar infarct size to that induced by permanent ischemia, that is,
penumbra does not exist after 4 hours. But in humans, some penumbra can still be detected
up to 48 hours from the beginning of symptoms.20, 21
Recent data suggest that not only penumbra, but also ischemic core is heterogeneous and
dynamic. It is hypothesized that, in the early minutes and hours of ischemia onset, the core
contains islets of injury (“mini-cores”), surrounded by salvageable viable tissue pockets
(“mini-penumbras”).13 Microvascular heterogeneous responses to ischemia support this
hypothesis.22 Saving the penumbra is the main target of acute stroke therapy, but this new
theory implies that with extremely early interventions we may have impact on progression of
infarct core as well.
Evidence on cell death mechanisms in the core is scarce, because if not all, most of the
agents failed to prevent damage in the core region. An exception is experimental use of
antioxidant uric acid,23 which is as effective as thrombolysis,24 and suggests that a major
12
mechanism of cell death in the core is generation of free radicals. In general necrosis
dominates in the core and apoptosis in the penumbra, but some ischemic cells may exhibit
combined biochemical features of apoptotic and necrotic pathways.25
1.1.1.2 Ischemic cascade
Leading pathogenic mechanisms of ischemic cascade include energy failure, elevation of
intracellular Ca2+ level, excitotoxicity, spreading depression, generation of free radicals,
blood-brain barrier (BBB) disruption, inflammation, and apoptosis (Figure 2). These follow
each other in a certain pattern, but not strictly in order, because they have overlapping
features. Progression of ischemic brain injury may last hours to days, inflammation and
apoptosis being the most long-lasting events.
ISCHEMIA
Energy failure
-acidosis
-Na+/K+ ATPase failure
-Extracellular K +
-Intracellular Ca2+
-Glutamate release
minutes
hours
Excitotoxicity
-free radical formation
-BBB damage
-inflammation
-lipid peroxidation
Necrosis
Apoptosis
days
weeks
Angiogenesis
Neurogenesis
Axonal remodeling
Figure 2 Mainstream events following focal cerebral ischemia.
First consequence of CBF reduction is the depletion of substrates, particularly oxygen and
glucose, which causes accumulation of lactate via anaerobic glycolysis. Acidosis potentiates
oxidative injury.26 With energy depletion membrane potential is lost and neurons and glia
depolarize.27 Energy failure leads to perturbation of the Na+/K+-ATPase and Ca2+/H-ATPase
13
pumps; in addition Na+-Ca2+ transporter is reversed.28 Excitatory amino acids (mainly
glutamate) are released into the extracellular space and intracellular levels of Na+, Ca2+, Cland extracellular level of K+ are increased. In contrast to core cells where anoxic
depolarization occur, peri-infarct penumbral cells can repolarize due to partly preserved
reperfusion, but depolarize again in response to increasing glutamate and K+ levels.
Spreading depression defines depolarization starting within the ischemic core and
extending outwards to surrounding tissue. It is an energy consuming process with the
duration and number of peri-infarct depolarizations being correlated with the final infarct
volume.29-31
Excitotoxicity of accumulated glutamate and imbalance of ions lead to activation of a
variety of Ca2+ dependent enzymes, including protein kinase C, phospholipase A2,
phospholipase C, cyclooxygenase, calcium-dependent nitric oxide synthase, calpain, various
proteases, and endonucleases. As a result, free-radical species and leukotrienes generate,
leading to irreversible mitochondrial damage, inflammation, and both necrotic and apoptotic
cell death. Due to the formation of mitochondrial permeability transition pore, the
mitochondrial membrane becomes leaky. This follows two important events: first, a burst of
free radicals32 and second, the release of cytochrome C.33 Free radicals react irreversibly
with several cellular constituents such as proteins, double bonds of phospholipids, and
nuclear DNA. Further, in conjunction with a weakened scavenger system, free radicals cause
lipid peroxidation, membrane damage, dysregulation of cellular processes, and mutations of
the genome. Cytochrome C is a central mediator of apoptosis (programmed cell death).
Other triggers of apoptosis include oxygen free radicals, death receptor ligation, DNA
damage, protease activation, and ionic imbalance. Pro-apoptotic signals lead to caspase
activation. Activated caspases are protein-cleaving enzymes, which lead to characteristic
DNA-laddering and cleavage of structural proteins (such as laminin, actin, gelsolin).
Apoptotic cell, differing greatly from necrotic cell, is characterized by: shrinkage of the
cytoplasm, marked condensation of chromatin, and fragmentation of the cell.34 Apoptotic
cells are rapidly removed by phagocytosis without eliciting an inflammatory reaction.35
An acute and prolonged inflammation reaction contributes to ischemic injury. Within minutes
of arterial occlusion, residents cells (mainly microglia) are activated and along with other
affected brain cells produce a plethora of proinflammatory mediators (tumor necrosis factorα,36 interleukin-6,37 monocyte chemoattractant protein-1,38 interleukin-1β,39 and granulocyte-
14
colony stimulating factor40). Especially monocyte chemoattractant protein-1 appears as a key
molecule in post-stroke inflammation that leads to transmigration of hematogenous
leukocytes.38, 41 After the expression of adhesion molecules (including intercellular adhesion
molecule 1 and selectins) at the vascular endothelium, neutrophils are the first inflammatory
cells to arrive to the ischemic tissue, as early as within hours after reperfusion, followed by
macrophages and monocytes within few days.42 Microvascular obstruction by neutrophils
(no-reflow phenomenon) can worsen the degree of ischemia, production of toxic mediators
by activated inflammatory cells and injured neurons can amplify tissue damage. Pathogenic
role of neutrophils and other leukocytes in cerebral ischemia is still a subject of debate.42
Inflammation seems to exert a dual role, acutely worsens the ischemic injury and in the longterm proves beneficial through tissue remodelling.43
Another important component of ischemic pathophysiology is edema formation. Edema is
initially cellular (cytotoxic edema) and follows cellular metabolic disturbances. Cytotoxic
edema is an important indicator of ultimate final infarct size.44 With the disruption of the
BBB, intravascular fluid leaks into brain tissue (vasogenic edema), increases tissue volume,
and proves number one reason of mortality in stroke. Mechanical or hypoxic damage of
vascular endothelium, toxic damage of inflammatory molecules and free radicals, and
especially destruction of the basal lamina by matrix metalloproteinases are potential causes
of BBB disruption. This issue will be further discussed in a dedicated section.
Late reperfusion can exacerbate the injury initially caused by ischemia. This so-called
reperfusion injury was first noticed when three hours of focal ischemia followed by three
hours of reperfusion in the rat have produced more damage than six hours of permanent
ischemia.45 Reperfusion injury triggers alterations in production of various cytotoxic
substances, including free radicals, excitatory amino acids, free fatty acids, proinflammatory
cytokines, and adhesion molecules.46 Involved pathological processes include leukocyte
infiltration, platelet and complement activation, postischemic hyperperfusion, and BBB
disruption.47 In some AIS patients, thrombolysis follows fatal edema or intracranial
hemorrhage; underlying event of these disastrous outcomes is reperfusion injury.
A late phase in the pathophysiology of infarcted tissue is tissue remodeling. Days to weeks
after a stroke, an active process of neural progenitor cell proliferation occurs. Neurogenesis
has been documented in several focal ischemia models.48 Trophic factors, such as brain-
15
derived neurotrophic factor and granulocyte-colony stimulating factor (G-CSF), promote
neurogenesis.49, 50 Active angiogenesis is observed both in experimental animals and
humans 3 days and further after an ischemic insult.51 Functional neuroimaging has
demonstrated changes in a number of features of brain function related to recovery after
stroke, including a period of functional growth characterized by demonstrable structural and
functional changes in both the ipsilateral and contralateral hemispheres that last several
weeks.52, 53
1.1.2
Acute ischemic stroke therapy
The main approach of treating AIS is timely recanalization. Thrombolytics or mechanical
devices (not yet evidence-based) are used to open the occluded artery and thus to restore
blood flow. In experimental scenario, a lot of of resources were spent on developing an
effective neuroprotective agent, which would protect neurons from necrosis, by interacting
with the components of ischemic cascade.
1.1.2.1 Recanalization
In 1996, intravenous recombinant t-PA became the first US Food and Drug Administrationapproved treatment for AIS. Original labeling requires the drug be given in 3 hours after
symptom onset based on the National Institute of Neurological Disorders and Stroke t-PA
trial.54 A restricted conditional license for the use of t-PA was granted in Europe in 2002
allowing treatment within 3 hours in AIS patients younger than 80 years who also met other
specified criteria. Results from the ECASS-III trial have since expanded time window to 4.5
hours.55 Thrombolysis with t-PA significantly improves clinical outcomes at 90 days
compared with placebo, for every 1000 patients treated, 97 more are alive and avoid
disability.56 The sooner the t-PA is given, the greater the benefit is, especially within 90 min
from symptom onset.57 A recent systematic review of 12 randomized trials of t-PA indicates
the same fact, the greatest benefit is gained with early treatment, although some patients
might benefit from t-PA up to 6 hours after stroke.56 Unfortunately, only a minority of AIS
patients receive t-PA: in the U.S. 3% to 8.5%58 and in the capital area of Finland as much as
15%59 of all AIS patients. The main reasons for this low rate of utilization of t-PA are
economical and educational: Cost of acute stroke care is high (though probably costeffective60) and awareness level on stroke in general public is low. The risk of symptomatic
hemorrhage (approximately 6%56) is another cause for withholding t-PA therapy.
16
Several novel thrombolytic agents are under progress. A recent successful example is
tenecteplase, which was more effective than t-PA in a phase II trial.61 Desmatoplase in
Acute Ischemic Stroke-3 (DIAS-3) and DIAS-4 phase III trials test the safety and efficacy of
desmatoplase given three- to nine hours after symptom onset in patients with proved
vascular occlusion.62
Bridging therapy of concomitant use of both iv and intraarterial t-PA aims to combine the
fast effect of iv use and higher recanalization rates of intraarterial use. Although yet an
investigational technique,63 a recent meta-analysis found the bridging therapy safe and
effective, suggesting that it may be the first choice of therapy in AIS patients with proven
arterial occlusion.64 This hypothesis is tested in an ongoing phase III trial.65
Endovascular revascularization is a local therapy of of therapy for AIS patients who have a
contraindication for systemic thrombolysis or who do not respond to it. Large vessel
occlusions are less prone to recanalization by t-PA.66 Clot retrieval with FDA-approved
MERCI® Retriever has been an evolving treatment option due to positive results of MERCI
trial.67 Pooled analysis of MERCI trial and the trial of combination of embolectomy and
thrombolysis (Multi MERCI) provides additional evidence that increasing rates of
recanalization improves clinical outcomes.68 Recently, mechanical embolectomy was not
found superior to standard care within 8 h of symptom onset in large-vessel, anterior
circulation strokes (MR-RESCUE).69
The Penumbra system was approved for use in the United States in 2008 for “the
revascularization of patients with acute ischemic stroke secondary to intracranial large vessel
occlusive disease (in the internal carotid, middle cerebral artery M1 and M2 segments,
basilar, and vertebral arteries) within 8 hours of symptom onset”. Penumbra system initiates
recanalization by in situ suction of the clot. According to the pivotal study70 and publication of
results from post-market experience71 the device seemed effective and safe. A phase IV
study testing t-PA-Penumbra System combined therapy over standard t-PA therapy in
anterior circulation strokes with a large clot is ongoing.72
Ultrasound-enhanced thrombolysis increases recanalization compared with t-PA.73 A
phase III trial in the USA (CLOTBUST-ER) is under progress.74
17
1.1.2.2 Neuroprotection
To year 2003 over 1000 agents were tested in animals, out of which 60% exerted
neuroprotection from focal ischemia.75 Nearly 200 clinical trials used candidate
neuroprotective agents in humans with no success. Potential reasons of this translation
failure have been extensively discussed elsewhere.76-81 Main reasons are the methodological
differences between preclinical and clinical studies, inadequate quality of animal testing and
the heterogeneity of stroke in humans compared to homogenous experimental strokes in
animals.82-84 After these many failures to prove a neuroprotective efficacy of drugs tested in
AIS patients, recently, the Evaluating Neuroprotection in Aneurysm Coiling Therapy (ENACT)
trial showed that neuroprotection worked in reducing ischemic infarcts after the endovascular
repair of an intracranial aneurysm.85 This study perhaps will alleviate the ignorance of
neuroprotection to treat AIS.
Besides others,84 a group of candidates holds hope for neuroprotection: 1) Hypothermia
combined with t-PA is tested for efficacy in a phase III trial86 and The European Hypothermia
trial, EuroHyp-1, is also a randomized multicenter study to assess efficacy and safety of
hypothermia in ischemic stroke patients,87) Free radical scavenging has become a
promising approach with proven efficacy of NXY-059 in the first clinical trial SAINT-I.88
Unfortunately SAINT-II was neutral,89 but indicated that treatment window should match the
preclinical data and patient selection should be improved.90 Currently another free radical
scavenger, ebselen, is tested in a phase III trial in patients with a cortical infarct.91 Drug will
be started within 24 hours of stroke and given during14 days. Edaravone was introduced as
the first free radical scavenger for the treatment of stroke92 and is widely used to treat acute
ischemic stroke in Japan, although evidence from large-scaled clinical trials are lacking, and
3) Minocycline, an antiinflammatory and antioxidant agent, was safe and effective in small
open-label studies,93 a phase IV study has been terminated, but yet results are missing.94
Animal data suggest that combining thrombolysis with neuroprotection may be superior
to thrombolysis alone and may extend the time window of thrombolysis.95 Erythropoietin was
harmful for this purpose,96 and currently uric acid97 and hypotermia98 are tested in phase III
clinical trials. A novel approach of neuroprotection has proven effective in primates by
selective inhibition of N-Methyl-D-aspartate receptor neurotoxicity.99 This high-quality
preclinical work is giving hope that neuroprotection may be feasible in AIS patients and
should be tested in a narrow population at first.100
18
For protecting neurons from dying, by only focusing on mechanisms of injury, we may have
been seeing only half of the picture.101 Brain’s response to stroke is complex and includes
multiple processes of endogenous repair and remodeling.51, 101, 102 It is suggested that
candidate drugs for AIS should possess regenerative mechanisms of action in addition to
neuroprotective actions, in order to achieve sustained neurological improvement.103 Thus,
another group of drugs are designed to enrich neurorecovery and remodeling of the infarcted
tissue. Among potential neurorestorative and neuroregenerative compounds, G-CSF has
been extensively studied104 and appeared effective in reducing infarct size and enhancing
functional recovery through its anti-apoptotic and neurogenesis inducing capacities.105, 104
Despite these encouraging preclinical results, a phase II trial (AXIS-II) of G-CSF in AIS
patient resumed negative.100 However, the drug awaits efficacy testing in chronic stroke
patients.106 Cerebrolysin, a peptide preparation which acts as a neurotrophic factor, promotes
neuroplasticity and neurogenesis in experimental models, but a phase III trial of cerebrolysin
plus t-PA in AIS patients resumed negative.107
1.1.3
Magnetic resonance imaging (MRI) in acute ischemic stroke
Recent developments in AIS imaging have revolutionized our approach to acute stroke.
Noncontrast computed tomography (CT), due to its wide accessibility, is most commonly
used diagnostic tool for acute stroke diagnostics. Noncontrast CT has been sufficient as an
imaging modality in thrombolysis candidates to receive tissue plasminogen activator in 3hour therapeutic window,54 but MRI may suit as a brain clock replacing the currently used
epidemiological time clock when deciding patient recruitment to thrombolysis.108 Magnetic
resonance is superior to CT to detect early ischemic changes109-112 and to detect involvement
of more than one third of the middle cerebral artery territory.113, 114 Although CT remains the
best method for detection of intracerebral hemorrhage, MRI with T2*can be equally
sensitive.14
MRI is noninvasive, patient compatible with only few contraindications, has relatively high
spatial and temporal resolution with superior signal contrast, and clinical availability is
constantly increasing. MRI may serve in AIS for several purposes: to ensure the diagnosis, to
give insights on etiology, and to guide the selection of therapeutic approach.
19
1.1.3.1 Lesion evaluation by MRI
Diffusion-weighted imaging (DWI) has revolutionized stroke research since this technique is
extremely sensitive to the hyperacute phase of brain ischemia. Another advantage of DWI in
stroke imaging is high sensitivity in detection of multiple acute ischemic lesions. 115-117
Perfusion-weighted imaging (PWI) is a semiquantitative method of evaluating brain
perfusion, which provides imaging and measuring blood flow at the capillary level.
Combination of DWI to PWI provides further insights on ischemic lesion evaluation. It is
generally accepted that the difference between PWI-based hypoperfusion volume and the
DWI-based lesion volume (the so-called mismatch) is operationally approximate the ischemic
penumbra.118, 119
It is thought that AIS patients with DWI-PWI mismatch may respond to thrombolysis beyond
the therapeutic window. Unfortunately, first-ever randomized controlled study in this context
(EPITHET)120 failed to prove effectiveness of t-PA in the 3- to 6 h treatment window
concerning the primary end point (geometric mean relative infarct growth), probably due to
spontaneously recovering subjects among mismatch patients.121 However, when a modified
method for calculating mismatch volume (co-registration method) is used, reanalysis of
EPITHET patients indicates attenuation in infarct growth.122 CT perfusion imaging-based
detection of penumbra has driven the patient selection in the positive phase II trial of
tenecteplase.61 A candidate group of patients to whom to apply mismatched-based
thrombolysis are wake-up stroke patients (an estimated 25% of AISs occur during sleep123125
).
There is a lack of consensus regarding which PWI-derived parameter best defines
hypoperfused region or predicts infarct growth.126-129 Other unsolved issues in MRI-based
penumbra imaging are the optimal method of postprocessing,129 the use of an automated
processing software,130 and method for delineation core and hypoperfusion. Optimization of
PWI/DWI mismatch clearly requires further studies.
A controversial issue is the clinical relevance of reversibility of DWI lesion.131 The definition of
PWI/DWI mismatch is questioned after the recognition that the PWI boundary includes a
region of benign oligemia and that a portion of the DWI core is potentially salvageable with
rapid reperfusion.132 But, sustained DWI reversal seems infrequent and rarely clinically
relevant by altering PWI/DWI mismatch.133
20
1.1.3.2 BBB disruption: contrast-enhanced imaging
A wide spectrum of appearances of contrast enhancement of the ischemic lesion on the CT
scans was noticed since late 70’s. Different patterns of enhancement (e.g. homogeneous or
heterogeneous) were observed at different degrees (e.g. minimal to marked), and in different
phases of the stroke,134-137 with a frequency varying from 26% to 95%.137 Enhancement was
seen as early as first day and as late as 9 months after the onset of symptoms, 136 but most
often at two to three weeks after stroke.137 Even with modern imaging techniques the
frequency of increased BBB permeability (BBBP) in AIS patients varies considerably (from
20 to 88%).138-140 A higher frequency is recognized with dynamic imaging methods using
quantitative approaches.
An association of CT contrast enhancement to hemorrhagic transformation of the infarction
was noted more than 30 years ago.141 In agreement with animal studies,142-145 early BBB
disruption detected by parenchymal enhancement in human ischemic stroke was found
highly specific for hemorrhagic transformation.146-148 With the increasing use of multimodal
MRI and CT, parameters related to BBB damage are being discovered and recent studies
often indicate an important role for BBBP imaging in prediction of hemorrhagic
transformation. Dynamic perfusion CT is increasingly used for this purpose,140, 149-151 though
there exists controversy about the optimal method of acquisition of the data (first-pass vs
delayed-acquisition). If the permeability of the BBB is not large enough for blood cellular
elements to pass, hemorrhage will not occur, but BBB leakage to much smaller molecules
such as albumin, may cause edema.152 Perfusion CT data analyzed with a modified Patlak
model was used to estimate BBBP for predicting malignant MCA and the need for
hemicraniectomy.152
Another imaging marker for BBB disruption was suggested as the so-called hyperintense
acute reperfusion marker. This imaging finding was observed only on non-contrast follow-up
examinations if gadolinium was administered during initial MRI and was defined as delayed
gadolinium enhancement in the cerebrospinal fluid153, 154 on fluid-attenuated inversion
recovery (FLAIR) images, indicating an early BBB disruption, increased risk of hemorrhagic
transformation, and poor outcome.148, 155, 156 However, hyperintense acute reperfusion marker
is a common finding in elderly stroke patients and is not necessarily associated with
hemorrhagic transformation.157
21
DCE-MRI based quantification of BBBP via a kinetic model is not standardized yet. Kassner
et al.139 applied Patlak model to DCE-MRI data for this purpose and evaluated BBBP in ten
AIS patients (who did not undergo thrombolysis). Increased permeability was found in
patients who developed hemorrhagic transformation.
T2*-based BBBP MRI,158-160 despite its semiquantitative nature, is a strong alternative to
DCE-MRI from a practical point-of-view, because perfusion MRI with T2*-WI is a part of
routine imaging in most stroke units. Among several candidate T2*-based measures, relative
recirculation, which identifies abnormalities in the contrast recirculation phase, provided
comparable data to DCE-MRI and proved predictive in identifying patients with AIS who will
proceed to hemorrhagic transformation.160
22
1.2
1.2.1
EXPERIMENTAL ISCHEMIC STROKE
Why animal modeling
Modeling of ischemic stroke serves for two main purposes: first, to disclose underlying
pathological mechanisms of focal cerebral ischemia and then based on these deciphered
mechanisms to develop novel therapies for stroke. Secondly, novel imaging techniques are
explored and applied in animal models before their introduction into clinical practice.
Thrombolysis161 and DWI162 are the mainstream examples of translational success from
laboratory to clinics.
Ischemic stroke is a very heterogeneous disease in human, varying in etiology, lesion
location, and size and is complicated by concomitant diseases. Age, sex, and amount of
collaterals are among several other variables affecting the outcome of ischemic stroke.
Therefore, very large group of patients are required in clinical trials of stroke. On the other
hand, in experimental stroke, a more homogenous disease is mimicked by strictly controlling
variables, thus, with a small group of animals statistical power can be already reached,
saving money and time. Rodents, less costly and more ethically acceptable than larger
animals, are most often utilized in experimental stroke research, because of several reasons:
the resemblance to humans in cerebrovascular anatomy, moderate size allowing easy
manipulations, low costs, the relative homogeneity within strains, and last but not least, the
accessibility for use by transgenic technology. In contrast to lissencephalic brains of rodents,
large animals have gyrencephalic brains and considerable amount of neocortex akin to
humans. Application of sophisticated methods such as evoked potential monitoring,
electroencephalography, and functional imaging is easier in larger animals. It is
recommended that positive rodent studies must be replicated in higher species before
proceeding to clinical trials.163 Even among primates there are considerable differences in
brain anatomy and vasculature, stroke in macaques may best represent human AIS.99
1.2.2
Major rodent models of ischemic stroke
There is a rich diversity of focal ischemia models, among which none is capable to mimic all
aspects of human stroke, but most appropriate model can be chosen to address a specific
23
question. Model selection is especially important in preclinical drug developmental studies.
Recommendations from the STAIR committee must be followed in designing such studies.163,
164
Each model grants superiorities and shortcomings (for reviews see related articles165-169).
Blocking of blood flow in the MCA is often aimed in animal models, because half of all
strokes occur in the territory of this artery.7 Thromboembolic models mimic closely the
disease from etiological aspect and they are suitable for testing thrombolytics. However, the
most commonly used method of inducing focal cerebral ischemia is intraluminal occlusion of
MCA by a surgical monofilament (suture model), which allows strict control on the timing of
the reperfusion. Other models include surgical occlusion of the distal or proximal MCA,
endothelin-1 induced ischemia, photothrombotic ischemia, and embolic models using
artificial materials as embolus. Models requiring craniectomy are complicated by both
negative side effects of exposing brain to the atmosphere and protective effect of
craniectomy from malignant MCA infarction.
Either permanent or transient ischemia is modeled. Permanent ischemic models represent
clinical situation of nearly half of the AIS patients. Others experience transient ischemia by
either spontaneous170 or therapeutic recanalization.171 In transient ischemia models, an
ischemia period of 90 to120 min is most often used, because it induces sustained infarcts.
Transient ischemia longer than 3 hours allows studying a specific aspect of the stroke,
reperfusion injury, but do not contain penumbra.172 A candidate neuroprotective compound
needs to be tested in both permanent and transient ischemia models, because of different
underlying mechanisms in each type of ischemia.163
1.2.2.1 Thromboembolic models
Thromboembolic models use two main strategies to induce stroke: clot embolism from an
extracranial artery and in situ clot formation in distal MCA. Originally autologous thrombi
were injected into extracranial arteries to reach the more distal intracranial arteries.173, 174 In
these earlier embolism models, infarcts induced by a shower of emboli were variable in size
and early spontaneous recanalization occurred.174, 175 Important parameters for inducing
consistent CBF decline and reproducible lesions are formation, composition, and final
localization of the emboli. By mechanically processing a preformed thrombus, autolysis
resistant fibrin-rich emboli may be achieved.176, 177 Others178 preferred injecting multiple fibrinrich autologous clots into the external carotid artery one after another leading to consistent
24
infarcts without spontaneous recanalization. Another method for increasing reproducibility of
thromboembolic infarcts is endovascular delivery of an intact fibrin-rich embolus into the
segment of the internal carotid artery near the origin of the MCA by intraarterial
catheterization.179, 180 In situ thromboembolic stroke model ensures exact localization of the
clot by microinjection of purified thrombin into the lumen of distal MCA.181 This model induces
reproducible cortical infarcts in mice and responds to t-PA treatment when t-PA is introduced
20 min after clot formation. In situ thrombus model avoids potential damaging effects of
intraarterial catheterization, but necessitates a craniectomy. Yet efforts are spent to improve
reproducibility in the thromboembolic model.182
Thromboembolic models responding to thrombolysis183-185 suit for testing new thrombolytics
and combination therapies of thrombolysis and neuroprotection for acute stroke.186
Preclinical drug testing may use rabbit embolic models of stroke187 secondary to a rodent
model. A large vessel thromboembolic occlusion model in rabbits may allow testing also
mechanical devices in combination to thrombolysis.188
1.2.2.2 Suture occlusion of the MCA
Originally introduced by Koizumi et al.189 this model includes insertion of a monofilament
suture into the internal carotid artery and advancing until it blocks blood flow to MCA. Suture
MCAO model induces MCA territory infarctions involving both frontoparietal cortex and
striatum with good reproducibility and reliability even among investigators of varying
experience.190 Reperfusion is easily achieved by retracting the suture. The model is suitable
to apply in the MRI scanner.191 Several modifications have been made to the initial model by
using differently coated sutures or external carotid artery insertion of the suture. 192, 193 Suture
diameter, type of coating of the suture (with silicone or poly-L-lysine), and insertion length of
the suture are among factors affecting the outcome. Size of the filament correlates well with
the size of the infarct.194, 195 Silicone-coated suture causes larger and more consistent infarcts
then uncoated suture induces.196, 197 The deeper the suture insertion is, the greater the
achieved lesion is.198, 199
Suture occlusion model carries some complications: Subarachnoid hemorrhage due to
mechanical vessel rupture and hyperthermia due to hypothalamic injury may occur. Silicone
coating of the suture and laser Doppler-flowmetry guidance may reduce the incidence of
subarachnoid hemorrhage.200 Spontaneous hyperthermia can be avoided by limiting
25
ischemia duration to 90 minutes or less201 or adjusting suture tip to a size that does not
occlude the hypothalamic artery.166
Intraluminal suture MCAO model was suggested suitable for neuroprotective drug studies
because a substantial penumbra exists within the first 60-90 min of injury.202 An MRI study
showed, however, that the mismatch volume is larger and persists longer in thromboembolic
model relative to permanent suture MCAO model in rats.203 Transient suture MCAO was also
compared to embolic model combined with t-PA treatment (thrombolysis model).204 Even
though infarct sizes were equal in these two models, perfusion recovers immediately and
completely in the suture model, but slowly and incompletely in the thrombolysis model. The
latter is associated with increased BBBP in the periphery of the infarct. Recently, serious
concerns raised on the use of transient MCAO occlusion with suture model in preclinical
studies.205 It has been argued that prompt recirculation achieved with this model is
uncommon in naturally occurring strokes and the model misleads clinicians in translating an
appropriately long time window for the agent under investigation.
1.2.2.3 Other models
Surgical methods aim to expose MCA by one of the several surgical approaches, among
which orbital route is less traumatic.206 While electrocauterization of a portion of MCA results
in permanent occlusion, the use of microclips and ligature snares allows reperfusion.207, 208
Originally induced by frontoparietal approach,209 distal MCAO at the rhinal fissure spares
lenticulostriate branches and leads only a restricted cortical infarction in rats. Tamura’s
subtemporal approach210 gained a greater acceptance because with Tamura’s method more
proximal regions of the MCA are accessible and infarcts are more reproducible. Not only the
site, but also the length of the ligated portion of MCAO affects the outcome.211 To overcome
variability in infarcts several modifications to surgical MCO have been proposed: tandem
occlusion of the distal MCA and ipsilateral common carotid artery,212 tandem occlusion of the
distal MCA and bilateral common carotid arteries (three vessel occlusion technique),207 and
distal MCAO with temporary clip compression of both common carotid arteries.213, 214 Orset et
al.181 have utilized temporal approach in their in situ MCA thrombosis model. Main handicap
of surgical methods is their invasiveness. Additionally, they require good surgical skills and in
transient surgical models recanalization is abrupt as in the intraluminal suture method. On
the other hand, the site of occlusion is well-controlled.
26
In models of non-clot embolus, many compounds or materials were used as artificial
emboli.215-219 Embolization with multiple microspheres187, 220-222 has been proposed to
simulate atheromata and fat embolization. It induces permanent ischemia, infarct maturation
is slower (i.e. penumbra persists longer223) than MCAO models, and infarcts are multifocal
and heterogeneous.224
In photothrombosis models, after systemic intravenous injection of a photoactive dye
(traditionally Rose Bengal), a cortical brain area is irradiated by a light beam at a specific
wavelength through the intact skull.225 Resulting perioxidative damage to the endothelium
causes platelet adhesion and aggregation in both pial and intraparenchymal vessels within
the irradiated area.226 Originally, this model induces cortical ischemic lesion with acute
severe endothelial cell damage, BBB disruption, and edema formation.227 Additionally,
ischemic lesion involves relatively restricted penumbral area, because of the associated end
arterial occlusion. However, a variation of the original phothrombotic model, ring model,228
successfully induces a penumbra-like lesion.229 Use of thinner ring irradiation may allow
achieving late spontaneous reperfusion and a morphological tissue recovery.230 Another
modification was made to occlude the MCA,231 simply MCA was exposed surgically and
irradiated after the injection of the photoactive dye. In this variant, although penumbral part of
the lesion is also minimal,232 a therapeutic effect of rtPA has been shown.233 Advantages of
the cortical phothrombotic models are control on the location and size of the lesion, and low
mortality.
Endothelin-1 (ET-1) is a potent vasoactive peptide, which produces a marked
vasoconstriction.234 Either by subtemporal approach235 or intracerebral injection technique,236
ET-1 application onto MCA provides significant decrease of CBF in the MCA territory,
resulting in an ischemic lesion pattern similar to that induced by direct surgical MCAO.236, 237
Direct cortical application of ET-1 induces a semicircular infarct involving all layers of the
neocortex.238 Endothelin-1 induced MCAO model mimics slow recanalization (lasting
approximately 4 hours) in a dose dependent manner of applied ET-1.237, 239 Therefore, less
control on the ischemia duration and intensity are disadvantages of the model. Furthermore,
it is not suitable for testing thrombolysis.
27
A number of animal models exist to study posterior circulation strokes;240 more or less
invasive embolic models were described.241, 242 A main handicap of modeling ischemic stroke
originating from posterior circulation is lesser reproducibility compared to MCAO models.163
1.2.3
Preconditioning
1.2.3.1 Ischemic tolerance
Introducing the brain a nonlethal insult allows it to be prepared for the next ischemic insult,
i.e. preconditions the brain. Preconditioning triggers a defense mechanism as a product of
genomic reprogramming, which renders brain tolerant to the final lethal injury. This new
defense program may attenuate almost all steps of ischemic cascade; additionally, innate
survival mechanisms and endogenous repair mechanisms are enhanced.243
Ischemic tolerance occurs in two temporally distinct windows: early tolerance can be
achieved within minutes, but wanes also rapidly, within hours; delayed tolerance develops in
hours and lasts for days. The main mechanism involved in early tolerance is adaptation of
membrane receptors, whereas gene activation with subsequent de novo protein synthesis
and genetic reprogramming dominates delayed tolerance.244 Data on early ischemic
tolerance in the brain are scarce,245-249 ischemic tolerance occurs in the brain predominantly
with the delayed pattern. Cross tolerance, in which one type of insult promotes protection
from a subsequent different type of insult, has been documented in the brain in many
variations.250 Hypoxic preconditioning is a well-known example of cross tolerance.
1.2.3.2 Hypoxic preconditioning
A brief exposure to systemic hypoxia (i.e. hypoxic preconditioning; HPC) prior to transient
MCAO reduces infarct volume, blood-brain barrier disruption, and leukocyte migration.251, 252
Hypoxic treatment was first time used as a preconditioning trigger in a hypoxia-ischemia
model in neonatal rats.253 Later, HPC was proven protective from both transient and
permanent focal cerebral ischemia.254, 255 Sublethal hypoxia (11% oxygen for two hours)
applied 48 h prior to transient focal cerebral ischemia protected nearly half of the tissue-atrisk from undergoing infarction in several strains of mice.254 Compared to permanent
ischemia alone, hypoxia applied 24 h prior to ischemia results in 30% smaller infarcts.255
Single applications of varying hypoxia durations (1, 3, or 6 hours) are similarly efficient, but
28
protective effect abolishes after 72 hours.255 Repetitive hypoxic treatment, however, may be
protective from focal ischemia up to 4 to 8 weeks.256, 257
The mechanisms of hypoxic preconditioning and ischemic tolerance are still being
elucidated. Hypoxia-induced tolerance in brain is not blocked by glutamate receptor
antagonists, but is blocked by inhibitors of RNA and protein synthesis.258 A key factor in HPC
is the hypoxia-inducible factor-1 (HIF-1). As early as 1 hour after hypoxia and maximally at 6
hours, expressions of many HIF-1-regulated genes are increased.259 Hypoxia stabilizes alpha
subunit of HIF-1, which enters the nucleus in a dimerized form and results in the induction of
HIF target genes. Several HIF target genes contribute to protection from ischemia,255, 258, 259
and their products are involved in a wide range of adaptive and pro-survival events, including
cellular metabolism, proliferation, vascularization, iron homeostasis, and glucose
metabolism.243, 260 However, in neuron-specific HIF-1alpha-deficient mice, protective effect of
hypoxia from subsequent focal ischemia was significantly attenuated, but not completely
abolished, suggesting that alternative mechanisms of neuroprotection are also implicated in
HPC.261 Recently microvascular sphingosine kinase activity was found as an important
trigger of hypoxic preconditioning.262 Blocking sphingosine kinase activity nullifies protective
effects of prior hypoxia from transient ischemia. Chemokine signaling seems to be another
critical mediator to the induction of hypoxic preconditioning-induced ischemic tolerance. Mice
that lacked monocyte chemoattractant protein-1 also lost the capacity to become ischemia
tolerant, although they received hypoxic treatment.251
1.2.4
Outcome measures
In contrast to human studies, where primary outcome is usually neurological improvement at
3 months, experimental stroke mostly focused on the acute phase of ischemia and the main
determinant of the outcome has been infarct volume within few days after the onset of
ischemia, mostly ignoring functional outcome.263 If the model includes mild ischemia, lesion
progression is slow, 264 or if a neuroprotective drug265 or a preconditioning regimen is tested,
therapeutic effect may be transient;266 therefore, long-term outcomes should also be
evaluated in such circumstances. Data from preclinical studies suggest a poor correlation
between pathologic and functional improvements.267 Despite the lack of infarct size
improvement, behavioral assessment might reveal effectiveness of a neuroprotective
drug.268, 269 Lack of correlation between different outcome measures indicates that
behavioral, neurological, and histological endpoints are conjointly necessary.
29
1.2.4.1 Infarct volume
Infarct volume is traditionally evaluated post mortem with hematoxylin-eosin staining (i.e. the
gold standard histological method) or with less costly 2,3,5-triphenyltetrazolium chloride
(TTC). Other staining methods (Cresyl violet270 or silver staining271) are also available to
delineate the extent of the ischemic lesion. Image analyzing systems allow manual,
semiautomated, or fully automated delineation of the lesion area,272 273 after which lesion
volume is calculated by multiplying with slice thickness. The larger the infarct, the more
pronounced the edema and enlargement of the injured tissue by edema results in
overestimation of the infarct volume. Thus, ischemic volume should be calculated with the
correction of edema,274 especially in models using proximal occlusion of the MCA for periods
longer than 60 min. Besides absolute infarct volumes, the percentage of the hemisphere
undergone infarction can be reported to facilitate comparison of the data from different
laboratories.
In vivo MRI enables monitoring lesion progression by repeated imaging. With DWI sequence
ischemic lesion can be identified as early as 3 min after the onset of ischemia191 and MRIbased lesion volume at 72 h correlates well with the TTC-based infarct volume.275
1.2.4.2 Neurological status
Motor deficits are relatively objective end points of a rat stroke model and can be evaluated
by a number of easy and quick methods.192, 211, 276 Tests to examine the effects of focal
ischemia on more refined sensorimotor functions include: limb placing, beam walking, grid
walking, rotarod, sticky label test, and staircase test.277 A number of cognitive tests are also
available, among which Morris water maze is the prototype.278 However, in the MCAO model,
spatial memory deficits seem minimal and watermaze impairments may be attributable to
sensory and motor deficits.279
Recently a tendency rose to use composite scores for functional evaluation in rodents. Such
an increasingly used evaluation is the modified neurological severity scores, which includes a
composite of motor (muscle status and abnormal movement), sensory (visual, tactile, and
proprioceptive), reflex, and balance tests.280 Despite the simplicity of administering the tasks,
deficits may be specific to a certain modality or function, and could be masked by the
composite score.281 In addition, the reflexes tested in the modified neurological severity
30
scores (pinna and startle reflexes) are irrelevant to the damage induced with MCAO.
Following points should be considered to ensure a successful functional evaluation after
stroke in a rodent model: 1) The chosen battery of functional assessments should be able to
detect even mild impairments, 2) it is important to obtain baseline data before experimental
manipulations, 3) for tasks that require pre-training, animals must be properly trained before
surgery, 4) experimentalist should be blind to treatment conditions to help eliminate bias.281
1.2.5
Sources of variability in experimental ischemic stroke
Age and sex of the animals should be considered. Rodent stroke studies mostly subjected
young males.263 Female rats compared to male rats sustain smaller infarcts after MCAO,
even in the presence of an additional pathology, such as diabetes and hypertension.282-285
This is due to the protective effect of estrogens,286 which is lost after ovariectomy.285 Stroke
pathophysiology differs between the aged and young rats. Ischemic stroke in aged rats are
associated with increased ischemia/reperfusion injury, earlier disruptions of the blood-brain
barrier, exacerbated neuronal degeneration, higher mortality, reduced functional outcome,
and reduced angiogenesis.287-290 Response to t-PA291 or to a neuroprotective agent may also
vary depending on the age of the animal.292, 293 Effectiveness of a neuroprotective agent in
aged animals50 illustrates a larger target population for such candidate drug.
Strain-dependent alterations in ischemia susceptibility are well-recognized.294-298 Fischer rats
are quite unsuitable for suture MCAO.295 Sprague-Dawley rats are most often used in stroke
research, but with very variable results.166 Strain of the rat may be a factor affecting the
outcome in preclinical drug studies.299-301 Some authors suggest Wistar Kyoto rat the best
choice, because it has a sustained vascular anatomy and its genetic relationship to the
spontaneously hypertensive stroke-prone strains makes Wistar Kyoto rat an ideal stepping
stone for later preclinical evaluations.166
The spontaneously hypertensive stroke-prone rats are species susceptible to develop
larger and much less variable infarcts following MCAO compared to other rat species.296, 302
In these rats, cortical infarcts and cerebral hemorrhages occur spontaneously, but
predominant lesions are small subcortical lesions, most probably with an initiating event of
BBB disruption rather than vasospasm, thrombosis, or ischemia.303 Therefore, spontaneously
hypertensive stroke-prone rats are good candidates for lacunar stroke modeling,304, 305 but
31
high mortality rate restricts their use in MCAO models.
Once a candidate stroke drug is proven efficient in otherwise healthy animals, next step is to
know whether it retains efficacy in the face of comorbidities, such as diabetes and
hypertension. Animal experiments of neuroprotection rarely involve testing in these
conditions,75 although, rodent models of both type 1 and type 2 diabetes306 and
hypertension307 are available. Type I diabetic rats subjected to thromboembolic ischemia
exhibited resistance to thrombolytic reperfusion, larger infarction volumes, increased
intracerebral hemorrhage,308 and higher BBB leakage.309 Type 2 diabetic rats showed a
defective angiogenesis after transient ischemia.310 Using these models one can decipher
how these comorbidities can influence the pathophysiology of stroke.
Several physiological parameters need to be monitored and regulated if they alternate
during an experiment of stroke. High blood glucose levels are not allowed because a body
of evidence suggest a role for hyperglycemia concerning deterioration in infarct size,
functional outcome, and blood-brain barrier damage.311-315 Body temperature is easily
monitored with a rectal probe. Animals should be kept normothermic for the reasons that
hyperthermia worsens the ischemic damage316 and hypothermia is neuroprotective.317
However, a set of animals should be used for testing whether the experimental treatment
itself is inducing hypo- or hyperthermia. Arterial blood pressure and gases should be
closely followed, particularly in deeply anesthetized animals.
Many of the commonly used anesthetics provide some degree of neuroprotection318 and
most of the volatile anesthetics trigger ischemic tolerance,319 for reasons why experimental
groups off animals should receive the same anesthetic and the possibility that anesthetic
may interfere with the effects of a candidate neuroprotectant should be considered.
32
1.3
1.3.1
BLOOD-BRAIN BARRIER
Structure and functions, neurovascular unit
The concept of the BBB date back to the late 18th century when Paul Ehrlich noted that an
intravenously injected dye leaked into all the organs except into the central nervous
system.320 The nature of the BBB was debated well into the 1960s.321 Current understanding
of the basic structure of the BBB is built on an electron microscopic discovery, that capillary
lumen bridged by tight junctions (TJ) form a continuous, impermeable membrane, which
forms the primary anatomical substrate of the BBB.322 The concept of the BBB has continued
to be refined over the past few decades. Currently, it is growingly recognized that, not only
cerebral microvascular endothelial cells, but multiple cells (such as glial cells, pericytes, and
neurons) constitute together with extracellular matrix a functional unit, “neurovascular unit”,
of which the integrity is essential to maintain homeostasis (Figure 3).323, 324 Concerted
synergism of the elements of the neurovascular unit gives rise to a BBB, which is simply
more than the sum of its parts. The human adult BBB has the same approximate surface
area as a tennis court, and a fifth of the cardiac output, that is, 1 to 1.5 L blood, passes over
it every minute at rest.325 Microvessels involve an estimated 95% of the total surface area of
the BBB. This barrier makes the brain practically inaccessible for lipid-insoluble compounds,
such as polar molecules and small ions, for which transport have to take place via carriermediated or vesicular mechanisms. Gases and small lipophilic molecules can diffuse through
the BBB.
Figure 3 Schematic diagram of
neurovascular unit that comprises
neurons, endothelial cells, astrocytes,
and pericytes. Basal membrane
surrounds endothelial cells and
pericytes.
33
1.3.1.1 Endothelial cells and pericytes
Endothelial cells of the BBB are distinguished from other endothelial cells by a number of
aspects: the presence of TJs,322 high number of mitochondria,326, 327 small number of
caveolae (membrane-bound vesicles),328 lack of fenestrations,329 minimal pinocytotic activity,
and near absence of vesicular transport.330 The transendothelial electrical resistance, which
restricts ion permeability, is in the range of 1000–5000 Ω cm2 in brain capillaries,331 more
than a hundred times higher than in noncerebral capillaries. Maturation of the BBB
necessitates endothelial cell expression of specific molecules (overviews exist332, 333).
Specific transport systems selectively expressed in the membranes of brain capillary
endothelial cells mediate the directed transport of essential nutrients into the central nervous
system or of toxic metabolites out of the central nervous system.332 Transendothelial
transport occurs, among many others, for hexoses (glucose, galactose), amino acids,
purines, and nucleosides. A receptor-mediated transport system resides in brain endothelial
cells for many substrates, including low-density lipoprotein, insulin, immunoglobulin G, and
transferrin. Active efflux pumps are also expressed in endothelial cells. Three classes of
transporters are implicated in the efflux of drugs from the brain:1) monocarboxylic acid
transporters, 2) organic ion transporters, and 3) multidrug resistance transporters (prototype
is P-glycoprotein).334 Enzymatic roles of the endothelial cells comprise another level of barrier
between cerebral circulation and brain (“metabolic BBB”). A well-known example of this
enzymatic barrier is DOPA-decarboxylase within the endothelial cells, which restricts the
transfer of dopamine from blood to brain.
Pericytes are located at the abluminal surface of the microvessels and encircle with their
processes 30 to 70% of the capillary wall.335 They are ensheathed by basal lamina, which
separates them from endothelium and astrocyte end-feet (Figure 3). There is approximately
one pericyte for every two to four endothelial cells. Pericytes are multifunctional in the brain
and they are required for both the stabilization and maturation of the capillary, as well as the
BBB.336 Pericytes-lacking mice develop perinatally brain edema and hemorrhages due to
increased BBBP,337, 338 of which one essential reason is deficient TJ formation. In mouse
brain during ischemia, pericytes contract and impair capillary flow.339 Additional roles are
suggested for pericytes in angiogenesis and neurogenesis occurring after stroke.336
34
1.3.1.2 Basal lamina
The basal lamina separates endothelial cells of brain vasculature from its neighboring cells
(Figure 3). It is composed of different extracellular matrix proteins, including collagen and
laminin. Matrix adhesion receptors, which are essential for the maintenance of the integrity of
the BBB, are expressed in the endothelial cells, neurons, and glia. Integrin and dystroglycan
receptors appear to bind endothelial cells and astrocyte end-feet to the individual intervening
matrix components.340
Focal ischemia initiates a rapid loss of integrity of the extracellular matrix within the
microvasculature and matrix adhesion receptors.340 With the disappearance of antigens of
the three main constituents of the basal lamina (laminin, fibronection, and collagen type IV),
the basal lamina loses its integrity.22, 341 Loss of the matrix proteins has been associated with
the rapid generation of members of four protease families: matrix metalloproteinases
(MMPs), serine proteases, cysteine proteinases, and heparinase, sources of which have not
entirely been worked out.342 Several lines of evidence from animal stroke experiments
suggest involvement of MMP-2 and MMP-9 in digestion of basal lamina leading to BBB
disruption, edema, and hemorrhagic transformation.341, 343-348 Additionally MMPs contribute to
the disruption of TJ proteins.349, 350 Caveolin-1 was recently discovered as an upstream
regulator of MMP activity after ischemia-reperfusion injury.351 A systematic review of AIS
patients indicated that serum MMP-9 levels are significantly correlated with infarct volume,
severity of stroke, and functional outcome, and MMP-9 may be a predictor of development of
intracerebral hemorrhage in patients treated with thrombolytic therapy.352
1.3.1.3 Tight junctions
TJs appear in endothelial and epithelial cells as a system of fusion with two main
parameters: the complexity of strands and the association of the particles with the inner (Pface) or outer (E-face) lipidic leaflet of the membrane. Brain endothelial tight junctions are the
most complex in the whole body vasculature, with respect to high number of strands (which
reflects high transcellular electrical resistance) and high P-face association. TJs, along with
adherens junctions, form a circumferential zipper-like structure between endothelial cells,
limiting paracellular passage of hydrophilic molecules. The degree of tightness of this zipper
varies within the microvasculature, as capillary endothelium proceeds to post-capillary
venous endothelium, strand complexity of TJs is reduced. The detailed molecular structure of
35
the TJs and the impact of ischemia on BBB with respect to TJs are reviewed elsewhere.353-357
Here, only main components of TJs are briefly summarized.
Junctional proteins can be categorized as transmembrane proteins and peripheral
membrane proteins. Transmembrane components of the TJ include junctional adhesion
molecule (JAM)-1, occludin, and claudins. Peripheral membrane proteins associate with TJs
in the cytoplasm; these are membrane-associated guanylate kinase –like proteins, including
zonula occludens (ZO)-1, ZO-2, and ZO-3.
Occludin, the first transmembrane TJ protein discovered,358 has four transmembrane
domains with two extracellular loops. Occludin is not mandatory for TJs or TJ strands to
form,359 however, presence of occludin is correlated with increased transcellular electrical
resistance and decreased paracellular activity.360 It is a critical regulatory protein for
mediating TJ responses in disease states.357 The carboxy-terminal of the occludin binds to
ZO, which in turn binds to the actin cytoskeleton, localizing it to the cellular membrane.
Dissociation of occludin from ZO may be related to increased BBBP after ischemia.361
The claudins, which share a similar membrane topography with occludin, but no sequence
homology, are believed to be the major transmembrane proteins of TJs, because occludin
knockout mice are still capable of forming these inter-endothelial connections, while claudin
knockout mice are nonviable362 and claudin-5 gene lacking mice show a loss of BBB
integrity.363 It is believed that claudins are responsible for the regulation of paracellular
permeability through the formation of paired strands.355 Among more than 20 identified
members of claudins, at least four (claudin-1, -3, -5, and -12) are expressed by BBB
endothelial cells, however, claudin-1 seems to be not targeted to the TJ.355 Claudins interact
directly with all ZO proteins. Claudin-5 and occluding mRNA expression are decreased and
these TJ proteins are degraded by MMP-2 and MMP-9 early after focal ischemia.349
JAMs are a family of immunoglobulin superfamily proteins that localize within the intercellular
cleft of TJs. JAMs participate in the assembly and maintenance of the TJs,357 overexpression
of JAMs in cells that do not normally form TJs increases their resistance to the diffusion of
soluble tracers, suggesting a role for permeability control for JAMs.364Among several JAMs
identified, JAM-A is highly expressed in the cerebrovasculature, but the status of JAMs after
stroke has not yet been studied.
36
ZO-1 was the first peripheral membrane component identified at TJs.365 Since then, many
further cytoplasmic components of TJs have been described, such as ZO-2, ZO-3, cingulin,
and afadin among others. Because the vast majority of experiments, addressing the role of
these proteins for TJ formation and regulation, were performed with epithelial cells, the BBB
related information on peripheral membrane proteins of TJs is still limited. ZO-1 acts as a
central organizer of the TJs, linking its carboxy-terminal region to the actin cytoskeleton.
Further, ZO-1 translocation from TJ membrane to cytoplasm is associated with an increased
barrier permeability.366 ZO-1 expression is reduced 24 h after focal cerebral ischemia and
this correlates with increased MMP-9 activity.346, 351, 367 MMP inhibition, nitric oxide synthase
inhibition, and knocking-out of MMP-9 gene, all prevent focal ischemia-induced ZO-1
degradation and BBB disruption.
1.3.1.4 Adherens junctions
Adherens junctions are primarily composed of vascular endothelial cadherin, which is
linked to cytoskeleton via catenins. The role of catenins in adherens junctions bears
resemblance to that of ZOI proteins in TJs.362 Disruption of adherens junctions at the BBB
can result in increased BBBP.321 Adherens junctions are functionally and structurally linked to
TJs, presumably playing an important role in the localization and the stabilization of the TJs
by forming a continuous belt localized near the apical end of the junctional cleft, just below
the TJs.368 The contribution of vascular endothelial cadherin and the catenins in BBB
disruption following stroke remains to be investigated.
1.3.1.5 Astrocytes
More than 99% of the surface of the brain capillaries is enveloped by astrocytic foot
processes, which allow communication between endothelial cells, neurons, and pericytes.
Many of the factors released by astrocytes (e.g. growth factors, cytokines, extracellular
matrix proteins) are able to induce specific features of the BBB in brain endothelium, which
are required during BBB development.366 Perivascular glial endfeet contribute to ionic, amino
acid, neurotransmitters, and water homeostasis at the BBB level.369 A close relationship
exists between the BBB and astrocyte membrane channel (aquaporin-4, AQP4). AQP4
dysregulation is coupled with BBB dysfunction and edema formation.370 However,
consequences of focal cerebral ischemia in AQP4 knockout mice are conflicting.371, 372 After
37
mild ischemia post-stroke brain swelling is not influenced by the lack of AQP4, but mice
score worse than their wild-type littermates.372 In contrast, after permanent ischemia swelling
and neurological scores are improved in AQP4 knockout mice.371
1.3.2
Methods to evaluate BBB permeability
The history of BBBP quantification stretches back at least 50 years.373 Currently, BBBP
studies utilize two main methods: in vitro analysis or in vivo imaging of an extravasated
exogenous tracer. Tracers can be classified into two categories: indicators for solute and ion
permeability and indicators for protein permeability.374 The most common contrast agents,
typically gadolinium diethylenetriaminepentaacetic acid (Gd-DTPA, MW 550 Da), are in vivo
imaging markers for solute and ion permeability, while Evans blue (EB) dye, which binds to
albumin (MW ≈68 kDa), is accepted as the gold standard marker for protein permeability.375
1.3.2.1 Qualitative methods
1.3.2.1.1 Visualization of dye extravasation
To monitor vascular protein leakage, in vivo studies most often used EB, which is an azo dye
(MW 980 Da) binding irreversibly to plasma albumin in a 10:1 molar ratio.376 The EB-albumin
complex extravasates from blood vessels into the surrounding tissues when the BBB is
disrupted. Intravenously injected 2% EB in saline (intraperitoneal administration is also
acceptable377) should circulate in the vasculature for a minimum of 30 min. Circulation times
vary from 20 min to 24 h between experiments, however, accumulation with longer
circulations than 30 min barely affects the results.377 Before terminating the experimental
animal, the dye is cleared from the bloodstream by transcardiac saline perfusion.
Macroscopically, blue-stained tissues indicate areas of BBB disruption. Red fluorescence of
EB (excitation at 620 nm, emission at 680 nm) can be visualized with a fluorescence
microscope or scanner.
1.3.2.1.2 Visualization of contrast agent extravasation
Intravenously injected contrast agent, of which bolus administration is preferable in
conditions other than tumors,378 leaks into extravascular space in the presence of BBB
disruption. Once in the extravascular space, voxels with higher concentrations of contrast
38
agent will appear bright on T1-weighted MR images due to T1 relaxation time shortening
caused by the contrast agent. Simple visual analysis of this enhancement has been found to
be a valuable tool to depict BBB disruption after many pathological conditions including AIS.
Early BBB disruption visualized with MRI predicted subsequent hemorrhagic transformation
in both experimental142, 143 and clinical settings.147, 379, 380 Although qualitative visual evidence
of parenchymal enhancement on postcontrast T1-weighted MRI is a highly specific
(specificity approximately 85%) predictor of hemorrhagic transformation, it is infrequent
during the crucial hours after symptom onset and insensitive (sensitivity near 35%).381
1.3.2.2 Quantitative methods
1.3.2.2.1 Colorimetric and fluorometric methods
The extravasated EB is extracted after the brain tissue is homogenized, centrifuged, and the
supernatant is diluted. Homogenization can include the entire brain, ischemic hemisphere, or
ischemic area only. Colorimetry at the absorbance of 600 to 620 nm after the subtraction of
background (baseline absorbance between 500 and 740 nm) determines EB content within
the limitations of the blue color. Fluorescence spectrophotometer at an excitation wavelength
of 620 nm and an emission wavelength of 680 nm detects EB 100 times more sensitively
than the colorimetric method.382, 383
1.3.2.2.2 Autoradiographic method
Although radiolabeled compounds enable highly sensitive measurements, safety concerns
and the need to process tissues for scintillation counting preclude immunochemical or
histological evaluation in the same experimental animals.384 Radiolabeled tracer is left in the
circulation for some time (typically 10 to 30 min), when a number of arterial blood samples
are collected. Brain samples undergo several procedures lasting over 24 h, thereafter Beta
counting is performed in a spectrometer. Blood concentration of the radiolabel is measured
by liquid scintillation counting. A blood-to-brain transfer ratio for the tracer is then determined
according to an analytic method.385 14C-sucrose (MW 342 Da), 3H-sucrose (MW 341 Da), 3Hinulin (MW 5 kDa), and 3H-aminoisobutyric acid (MW 103 Da) were often applied in
experimental models of stroke studying BBBP.45, 344, 386, 387
39
1.3.2.2.3 Fluorescence methods
Fluorescent-labeled tracers can be introduced intravenously; after obtaining simultaneous
blood samples, tissue and blood concentration of fluorescent-labeled tracer can be quantified
by a spectrofluorometer. By analyzing these data with an analytic method, as it is done in
autoradiography analyses, BBBP to the tracer is estimated. Availability of intravital confocal
microscopy to monitor extravasated fluorescent tracer is a substantial advantage of
fluorescent-labeling technique.388, 389 BBBP to both proteins and solutes can be examined via
the use of fluorescein isothiocynate-albumin (MW 67 kDa)390 and sodium fluorescein (MW
376 Da),384 respectively.
1.3.2.2.4 Other methods
Immunohistochemical staining methods enable the detection and quantification of an
extravasated endogenous macromolecule, such as IgG389 and albumin,391 or of blood cells,
such as polymorphonuclear leukocytes.389 A recent technique, near-infrared optical imaging,
provides intravital evaluation of BBBP with a higher spatial resolution than intravital confocal
microscopy.392, 393
1.3.2.2.5 Dynamic contrast-enhanced MRI (DCE-MRI)
As mentioned previously, contrast enhancement of the brain tissue indicates BBB leakage.
Dynamic method of permeability assessment is more sensitive to subtle T1 enhancements
than simple postcontrast SE imaging alone, because DCE-MRI provides about several
dozens of sampling times to characterize the enhancement, which increases conspicuity of
low “effect-to-noise” ratio.139
Determination of contrast agent concentration is a challenge that usually is solved via the
measurement of the T1 of the tissue and the change in T1, with the assumption that the
increase in T1 relaxation rate is proportional to the concentration of contrast agent.394 For this
purpose, before, during, and after an intravenous bolus contrast agent, T1-weighted
gradient-recalled echo imaging of the brain is repeated dozens of times over a course of
several minutes. Then, one can analyze the data gathered with this DCE-MRI technique by
one of available mathematical models,394 quantify contrast as a function of time, and estimate
BBBP.381 In stroke animal experiments, a graphical approach to analyze DCE-MRI data, the
Patlak plot approach,395 proved relevant in comparison to gold standard methods.142, 396, 397
40
and is increasingly used.291, 389 Clinical use of BBBP imaging with DCE-MRI and Patlak
method is yet in its infancy, but retains promise for estimating clinical prediction of
hemorrhagic transformation after t-PA treatment.398 A disadvantage of the BBBP MRI is the
long acquisition time of the images, however, a recent study reported three and a half
minutes as feasible.399
1.3.3
BBB disruption in experimental stroke
The profile of BBB disruption following transient focal cerebral ischemia is typically described
as biphasic, referring constantly to same studies (more specifically to studies by Kuroiwa et
al,400 Belayev et al,401 Rosenberg et al,344 and Huang et al386). These deserve a closer look
and will be discussed in detail in Discussion. Table 1 resumes characteristic features of all
animal studies which reported biphasic BBB disruption after focal ischemia-reperfusion
injury. If one attempts to combine results from all these studies (Figure 4), it becomes clear
that some agreement exists on the decreased leakage around 24 h and increased leakage
around 48 h after reperfusion.
Table 1 Animal studies of transient focal ischemia reporting biphasic BBB leakage
Ischemia model
Animal
BBB evaluation
method
Evaluated
time-points*
Open
Closed
Less open More open
References
400
Proximal MCA ligation
cat
EB
2, 3, 5, 72 h
0-2 h
3h
NA
5, 72 h Kuroiwa et al.
for 1 h
qualitative
Intraluminal MCAO
rat
EB
0-1, 1-2, 1-3,
1-3 h
0-1 h
22-24 h
46-48 h Belayev et al.401
for 2 h
quantitative
22-24, 46-48 h
Intraluminal MCAO
rat
14C-sucrose
3, 6, 15, 24,
3h
336 h
6, 15, 24,
48 h
Rosenberg et al.344
for 2 h
quantitative
48, 120, 336 h
120
Distal MCA ligation SHR
3H-sucrose
1.5-2 min, 1, 4, 22, 46 h 1.5-2 min
NA
1, 4, 22 h
46 h
Huang et al.386
for 2 h
quantitative
Intraluminal MCAO
rat
EB
2, 6, 24, 48 h
6h
NA
2, 24 h
48 h
Wu et al.443
for 2 h
qualitative
Intraluminal MCAO
rat
CE-MRI
15 min, 3, 6, 24, 72 h 15 min-6 h
NA
NA
72 h
Veltkamp et al.438
for 2 h
semiquantitative
Intraluminal MCAO
rat
CE-MRI
4, 24, 48 h
4h
NA
24 h
48 h
Pillai et al.444
for 1,5 h
semiquantitative
392
Intraluminal MCAO mouse
EB
0-4, 4-8, 8-12,
4-8 h
8-12 h
16-20 h
12-16 h Klohs et al.
for 1 h
qualitative
12-16, 16-20 h
Intraluminal MCAO mouse
NIRF imaging
4-8, 8-12, 12-16 h
4-8 h
8-12 h
12-16 h
NA
Klohs et al.392
for 1 h
qualitative
*time-points after reperfusion
BBB, blood-brain barrier; EB, Evans blue; CE-MRI, contrast-enhanced magnetic resonance imaging; MCA, middle cerebral artery; MCAO, MCA
occlusion; NA, not applicable; NIRF, near-infrared fluorescence; SHR, spontaneously hypertensive rat
41
Figure 4 Compiled BBB opening data of the studies resumed in Table 1. The ordinate is the
BBB status. Closed circle indicates data collection at a single time-point; arrow indicates data
collection during a period. Note accumulation of the data on specific time-points (ovals).
1.3.3.1 Theories of biphasic BBB disruption
The first phase of the biphasic permeability has been attributed to increased inflammatory
and oxidative stress on the BBB, in conjunction with enzymatic degradation of the ECM by
MMPs.44, 344 A synthetic MMP inhibitor blocked this initial opening, but had no effect on
delayed aggravation of BBB leakage.344, 349 Associates of the final phase of the biphasic
BBBP appears as angiogenesis357 and induction of MMP-3 and MMP-9 by cyclooxygenase2402, 403 The period between the two barrier openings, the “refractory period”, was assumed
as a result of normalization of the BBB function due to subsiding hyperemia and reestablishment of autoregulation.400 Alternatively, post-ischemic inhibition of pinocytotic
transport400 and no-reflow phenomenon were used to explain the transient “closure” of the
BBB.
Some authors acknowledge a triphasic course for BBB opening, adding a hyperacute
42
opening, which occurs immediately after reperfusion and prior to above stated phases.357
Theoretical explanation for this initial phase is opening of the BBB with passive disassembly
of TJs due to hyperemia and closing by the following pericyte contraction, which supports the
BBB.357 Actually, hyperemia has coincided with the first phase of opening reported by
Kuroiwa et al.400 and Huang et al. confirmed a hyperacute opening, but found as the first
phase of the biphasic opening.
1.3.3.2 Continuous BBB disruption
In recent years, with the increasing use of BBBP imaging in experimental stroke studies,
accumulating data suggest that BBB leakage to contrast agents (because most commonly
Gd-DTPA is used), thus to solute and ions, occurs in a continuous pattern. In addition, a
review of previous studies reporting a biphasic BBB opening (Table 1) disclose that some
studies in fact have failed to show any closure between two phases of increased BBB
leakage, therefore they are also included in Table 2, which summarizes animal studies
showing continuous BBB leakage after focal ischemia-reperfusion injury.
Table 2 Animal studies of focal cerebral ischemia suggesting a continuous BBB disruption.
Ischemia model
Animal
BBB evaluation method
Evaluated time-points*
Intraluminal MCAO
for 2 h
rat
albumin fluorescence
qualitative
6, 12 , 24 h
References
Albayrak et al.400
Proximal MCA ligation
for 3 h
rat
CE-MRI
qualitative
0.5, 1.5, 2.5, 3.5, 4.5, 6 h
Intraluminal MCAO
for 1 and 2.5 h
rat
CE-MRI
qualitative
0.5, 1.5, 2.5 h
Neumann-Haefelin et al.434
Intraluminal MCAO
for 2 h
rat
EB
qualitative
2, 6, 24, 48 h
Wu et al.
Intraluminal MCAO
for 1.5 h
rat
CE-MRI
semiquantitative
0, 1, 4 days
Lennmyr et al.435
Intraluminal MCAO
for 2 h
rat
CE-MRI
semiquantitative
3, 5, 8, 12 h
Nagel et al.436
Intraluminal MCAO
for 2 h
rat
CE-MRI
semiquantitative analysis
15 min, 3, 6, 24, 72 h
Intraluminal MCAO
for 1.5 h
rat
CE-MRI
semiquantitative analysis
3.5, 24, 120 h
Three-vessel occlusion
for 1 h
rat
DCE-MRI
quantitative
1, 3, 7, 14, 21 days
Kastrup et al.433
443
Veltkamp et al.438
Nagel et al.437
Lin et al.
439
Pillai et al.444
Intraluminal MCAO
rat
CE-MRI
4, 28, 48 h
for 1.5 h
semiquantitative analysis
*time-points after reperfusion
BBB, blood-brain barrier; CE-MRI, contrast-enhanced magnetic resonance imaging; MCA, middle cerebral artery; MCAO, MCA occlusion
43
1.4
BRAIN ISCHEMIA AND STANNIOCALCIN
Stanniocalcin (STC) is a glycoprotein originally discovered in bony fish, in which it acts as a
calcium/phosphate regulating hormone.404 In human, two members of STC homologs were
discovered: STC1, which shares about 72% amino acid sequence and 80% protein similarity
to fish405 and STC2, with approximately 34% identity and 60% homology to amino acid
sequences to STC1.406 STCs are made in virtually all tissues and regulate various biological
functions. Based on their ubiquitous expression patterns and generally undetectable levels in
blood serum, it is unlikely that the mammalian STCs play important roles in serum Ca2+/Pi
homeostasis.407 Accumulating evidence denote the involvement of STC1 and STC2 in the
subcellular functions of mitochondria and endoplasmic reticulum, respectively, particularly
responding to oxidative stress and unfolded protein response.407-411
STCs are widely expressed in brain neurons, in choroid plexus epithelium, and to some
degree in microvascular endothelial cells.412-415 Considerable data suggest a neuroprotective
role for STCs. Pathological events, such as focal ischemia,416 traumatic injury,417 focal
cryoinjury,418 and hypoxia418 increase brain levels of STC1. Overexpression of STC1 was
linked to expression of interleukin 6 (IL-6)418 and increased neuronal resistance to hypoxia
and hypercalcemia in vivo.416 Enhanced brain expression of STC2 followed focal brain
ischemia and hypoxia.411 Intracerebroventricular injection of STC2 protected hippocampal
neurons from kainic acid toxicity.419
There exist no previous data regarding STCs and the BBB. Some STC1 staining was
observed in brain endothelial cells lining the capillaries.413 In studies of angiogenesis, STC1
upregulation is a striking finding in endothelial cells.420 A role for STC1 in maintaining
permeability in human coronary artery endothelial cells emerged during cardiovascular
inflammation.421 In vivo hypoxia induced, among others, STC genes in vascular endothelial
cells.422 In human brain microvascular endothelial cells, β-amyloid peptide toxicity induces
considerable STC-1 gene expression, which possibly exerted some antiinflammatory and
antiapoptotic effects.414
44
2 AIMS OF THE STUDY
1. To critically address the presumed biphasic BBB opening following transient focal cerebral
ischemia by a systematic, quantitative, and multimodal approach. (I)
2. To compare the gold standard ex vivo method of BBBP evaluation (quantification of
extravasated Evans blue dye) to an in vivo method (DCE-MRI), that may be applicable to
humans. (I,II)
3. To compare BBBP to a large (Evans blue-albumin, MW ≈68 kDa) and small molecule
(contrast agent Gd-DTPA, MW 550 Da). (I,II)
4. To characterize spatial features of BBB leakage within the ischemic area and to find any
correlation between the extent or severity of ischemia and the degree of BBB leakage. (I,II)
5. To extend the findings from studies I and II with a more powerful study design, that
included repeated evaluations of post-ischemic BBBP in the same animals. (III)
6. Using genetically modified STC-1 deficient mice in the scenario of transient focal ischemia
with and without prior HPC, to test the necessity of STC-1 for HPC to occur; additionally, to
uncover the effect of STC-1 deficiency on BBBP. (IV)
45
3 MATERIALS AND METHODS
All experiments were carried out in Biomedicum Helsinki on the premises of Department of
Neurology, Helsinki University Central Hospital. The Animal Research Committee approved
all studies.
3.1
ANIMALS
Adult male Wistar rats (Harlan Nederland, Horst, The Netherlands), weighing 290 to 340 g
(I,II, and III) and adult male STC1 knockout (STC1-/-) mice on a C57BL/6 background
(courtesy of Dr. Andy C.-M. Chang, Children's Medical Research Institute, Parkville,
Australia) with NMRI (Harlan Nederland, Horst, The Netherlands) wild-type controls weighing
25 to 40g (IV) were used. Previous characterization of the STC1-/- mice revealed normal
viability and fertility and no gross histological or morphological abnormality.423 Prior to
surgery, rats were housed in groups of five and mice in groups of 10 and after the surgery
individually, in a temperature- and humidity-controlled environment and 12/12 h light/dark
cycle with free access to food and water. All efforts were made to minimize animal distress
and to reduce the number of animals used.
3.2
ANESTHESIA
Rats received an intraperitoneal injection of ketamine hydrochloride (50 mg/kg, Ketalar,
Parke-Davis, Detroit, MI, USA) and a subcutaneous injection of medetomidine
hydrochloride (0.5 mg/kg, Domitor, Orion, Espoo, Finland). Mice received same drugs, but
subcutaneously.
3.3
MONITORING OF PHYSIOLOGICAL PARAMETERS
In rats, arterial blood pressure monitoring (Olli Blood Pressure Meter 533, Kone, Espoo,
Finland) required left femoral artery catheterization by a PE-50 polyethylene tube. Left
femoral vein catheter served for injections of contrast agent Gd-DTPA (1mL/kg, Magnevist,
0.5 mmol/mL, Schering, Germany) (I, II, and III) and EB dye (3 mL/kg of 2% solution; 20
46
mg/mL dissolved in 1% albumin, Sigma-Aldrich, Steinheim, Germany) (I). Catheters were
removed after reperfusion, except in Study III the femoral vein catheter was kept in place for
one week. Rectal temperature was monitored and maintained at the physiological ranges
with a heating blanket during the surgery and MRI.
3.4
STUDY DESIGNS
Figure 5 Schematic presentation of study designs of Study I, II, and III. Rats (overall 123)
with adequate ischemia/reperfusion were allocated to 15 study groups based on the postreperfusion interval (I,II). Study III included repeated MRI of the same animals at 5 timepoints after reperfusion and MRI protocol was identical to that of Study II, except it included
additionally pre- and postcontrast FLAIR images.
DWI, diffusion-weighted imaging; FLAIR, fluid-attenuated inversion-recovery; Gd-DTPA,
gadolinium diethylenetriaminepentaacetic acid; IR-FLASH, inversion recovery snapshot-fast
low-angled shot; LDF, laser-Doppler flowmetry; tMCAO, transient middle cerebral artery
occlusion.
47
Figure 6 Schematic presentation of study design of Study IV. Each row of the schema
resumes a separate experiment, which included two groups of mice: Wild-type and
stanniocalcin knockout mice. Coll, collection; EB inj, Evans blue injection; HPC, hypoxic
preconditioning; Neurol eva, neurological evaluation; tMCAO, transient middle cerebral artery
occlusion.
3.5
FOCAL CEREBRAL ISCHEMIA MODEL
Intraluminal suture occlusion of the right MCA induced focal cerebral ischemia.274 Rat
occluders were prepared from 4-0 nylon monofilament suture (Ethilon Nylon Suture,
ETHICON Inc., Somerville, NJ, USA) with its tip rounded by heating near a soldering iron and
then coated with low viscous silicone (Provil L, Bayer Dental, Leverkusen, Germany) to a
0.38 to 0.40 mm diameter. Under general anesthesia and the rat lying on its back, the
common carotid artery and external carotid artery were exposed through a ventral midline
neck incision. The proximal common carotid artery and the origin of the external carotid
48
artery were ligated. The occluder was inserted into the common carotid artery via an
arteriotomy and was advanced into the internal carotid artery approximately 17 mm above
the carotid artery bifurcation. At this point, a mild resistance indicated that the tip of the
occluder was lodged into the proximal anterior cerebral artery and occluded the origins of the
MCA and the posterior communicating artery.
Mouse model was slightly different.424 Mouse occluders were prepared from 5.0
monofilament nylon suture (Ethilon Nylon Suture, ETHICON Inc., Somerville, NJ, USA) with
their tips rounded to 0.15 to 0.20 mm diameter. Occluder was inserted into the internal
carotid artery through an arteriotomy made in the external carotid artery. Insertion length
from common carotid artery bifurcation was approximately 9 mm. The ligature of the common
carotid artery was released after MCAO. Reperfusion was accomplished by withdrawing the
suture occluder 90 min (I, II, and III)
or 60 min (IV) after MCAO. Sham-operated animals underwent the same surgery, except
that the suture occluder was withdrawn before causing MCAO.
3.6
HYPOXIC PRECONDITIONING
Mice (6-8 per procedure) received hypoxia in an airtight transparent low-oxygenation
chamber perfused with 8% vol/vol oxygen in nitrogen for six hours.418 During hypoxic
treatment mice exhibited reduced locomotor activity, which was normalized within minutes
after returning to normal environment.
3.7
LASER-DOPPLER FLOWMETRY
CBF was measured in all animals with LDF (OxyFlo, Oxford Optronix Instruments, Oxford,
UK) by applying a flexible fiber-optic probe. The scalp was incised in the midline exposing
the skull. In rats, the skull area over the ipsilateral MCA region (1.0-2.5 mm caudal and 6.0
mm lateral from the bregma) was thinned by a dental drill to allow measurement via the
probe. In mice, without drilling, the probe was located over the territory supplied by the MCA
(2 mm caudal and 3-4 mm lateral from the bregma). CBF was repeatedly measured before
and during MCAO as well as after reperfusion.
49
3.8
MRI STUDIES
MRI studies were performed with a 4.7 T scanner (PharmaScan, Bruker BioSpin, Ettlingen,
Germany) using a 90-mm shielded gradient capable of producing a maximum gradient
amplitude of 300 mT/m with an 80-μs rise time. The linear birdcage RF coil used had an
inner diameter of 38 mm. Following shimming, a pilot imaging sequence (a multi-slice spinecho pulse sequence with repetition time/echo time: 200/8.9 ms, matrix size: 128Χ128, fieldof-view: 5.0 cm, number of averages: 1, slice thickness: 2 mm) served for reproducible
positioning of the animal in the magnet at different MRI sessions. DWI scans were acquired
using a spin-echo echo-planar imaging sequence (repetition time/ echo time: 4000/80 ms,
matrix size: 128x128, field-of-view: 40x40 mm, slice thickness: 2 mm) with three b values
(b0: 0.4, b1: 1280, and b2: 2342 s/mm2, diffusion was measured in the read gradient
direction). Longitudinal relaxation time (T1 value) measurements were obtained with an
inversion recovery snapshot-fast low-angled shot (IR-FLASH) sequence (repetition time/
echo time: 2.2/1.4 ms, 12 inversion delays from 140 to 3230 ms, flip angle: 5°, matrix size:
128×128, field-of-view: 40×40 mm, slice thickness: 2 mm, number of averages: 15). FLAIR
images were acquired with rapid acquisition with relaxation enhancement sequence
(repetition time/echo time: 10,000/38.6 ms, inversion time: 1800 ms, matrix size: 256×128,
zerofilled to 256×256, field-of-view: 40×40 mm, echo train length: 16, number of averages: 1,
slice thickness: 2 mm).
During imaging, rectal temperature was maintained at the physiological ranges by use of a
MRI compatible heating pad and pump (Gaymar Industries, Orchard Park, NY, USA).
During DWI, a 7-slice data set was obtained covering the entire brain except the olfactory
bulb. The first axial slice was selected posterior to the olfactory bulb, navigating by the rhinal
sulcus detected on the pilot image, and the following slices were placed caudally at 2-mm
intervals. Images of IR-FLASH and FLAIR were obtained with a single axial slice at the
coordinate of the third slice of the DWI, the optic chiasmal slice, which is approximately at 0.5
mm posterior to the bregma. All the analyses, except ischemic volume calculation, were
performed on this slice. The IR-FLASH scan with inversion delay −1826 ms was chosen as
the T1-weighted image (T1-WI). The DWI scan with b0 provided the T2-weighted image (T2WI). T1 maps were constructed from each IR-FLASH sequence and apparent diffusion
coefficient (ADC) maps from DWI sequences using ParaVision 2.1.1. Software (Bruker
BioSpin, Ettlingen, Germany). All image analyses described below used ParaVision 2.1.1.
50
Software. Regions of interest (ROIs) were placed manually on the ipsilateral hemisphere and
control ROIs were placed on the homologous locations in the contralateral hemisphere
(Figure 7).
Figure 7 Regions of interest (ROIs). A, from Study II and B, from Study III.
3.8.1
Patlak plotting
Patlak plotting395 is a graphical analysis method of the plasma and tissue MRI data, to
estimate the blood-to-brain transfer rate constant of the contrast agent (Ki) in the
permeability-limited circumstances of a two-compartmental model, such as focal brain
ischemia. It is assumed that there is a steady phase in the blood-to-brain distribution of the
tracer during which the tracer crosses the BBB in one-way direction, towards the brain tissue.
If plasma and tissue MRI data, which are collected repeatedly overtime, are plotted, this
results in an uptake curve with a linear phase of which the slope approximates Ki.
To make the Patlak plots, arterial plasma and tissue concentration of contrast agent (Gd-
51
DTPA) are needed. Intravenous administration of Gd-DTPA leads to a change in reverse
longitudinal relaxation rates (R1) of protons in its distribution area. With the assumptions that
1) the increase in relaxation rate (∆R1) is proportional to the concentration of contrast
agent394 and 2) tissue relaxivity (r1t) and plasma relaxivity (r1p) are the same,327 following
equation is acquired:
 R1t(t) = R1t(t)-R1t0(t) = r1tCt(t) and  R1p(t) = R1p(t)-R1p0(t) = r1pCp(t) (1-Hct)
where (t) is the duration of the experiment, R1t0(t) is the tissue longitudinal relaxation rate
before Gd-DTPA injection and R1t(t) at the end of experiment; R1p0(t) is the plasma
longitudinal relaxation rate prior to Gd-DTPA injection, and R1p(t), at the end of the
experiment; Ct(t) is the tissue concentration of the Gd- DTPA at the end of the experiment;
Cp(t) is the plasma concentration of GD-DTPA at the end of the experiment; Hct is the
hematocrit (arbitrarily 43%). Cp of Gd-DTPA was measured from the superior sagittal sinus.
This approach of approximating arterial input function of the tracer in arterial blood through
measurements from venous system has no significant affect on Ki results.425
Accordingly, the relation between tissue and plasma concentrations of Gd-DTPA is described
in the following equation:
t
Ct(t) = Vp Cp (t) + Ki

Cp (τ) dτ
0
where Cp (T) is the plasma concentration at a series of times over the duration of the
experiment and is used to calculate the arterial-concentration time integral; Ki is the blood-tobrain transfer rate constant of Gd-DTPA; Vp is the blood plasma volume.
t
Patlak plots are constructed by plotting Ct(t) / Cp(t) (ordinate) versus

Cp (τ) dτ / Cp (t)
0
(abscissa). The abscissa has the units of time, which is not the real time but the
concentration-adjusted time (tstretch). (Figure 8)
52
Figure 8 Patlak plots of a representative rat. Imaging was performed at 2 hours after
reperfusion that followed 90-min ischemia. The ordinate is the ratio of brain tissue
concentration of Gd-DTPA to its plasma concentration. The abscissa represents
concentration-adjusted time, stretch time. Plotted data showed linearity during whole imaging
time (20 to 30 min).
3.8.2
Imaging protocol
All rats underwent DWI to ensure the presence of stroke immediately after induction of
MCAO confirmed by LDF. Sham animals underwent MRI 24 h after sham operation,
otherwise MRI was run at the corresponding time-points after reperfusion (see Study design).
MRI protocol included: a pilot sequence, DWI, a pair of IR-FLASH sequence, and a postcontrast IR-FLASH sequence 25 min following a bolus of Gd-DTPA injection (I) or
repeated IR-FLASH sequences at approximately 1-min intervals for 20 to 30 min after GdDTPA injection (II, III) and pre- and post-contrast FLAIR sequences (III). MR imaging took 25
to 35 min.
53
3.9
NEUROLOGICAL EVALUATION
Sensorimotor performance of mice was scored 24 hours after reperfusion (IV), as follows:
0, normal; 1, contralateral paw paralysis; 2, contralateral paw paralysis plus decreased
resistance to lateral push; 3, 2 plus circulating behavior; 4, no spontaneous walking with
depressed consciousness; and 5, death.426
3.10 TISSUE HANDLING
After various periods of time following reperfusion (see Study design), animals were
reanesthetized with a lethal dose of intraperitoneal pentobarbital (1 mL/rat, 0.04 mL/mouse,
Mebunat, 60 mg/mL, Orion). For cardiac perfusion, after a midline abdominal incision, the
chest was opened. A catheter was inserted into the aorta via the left ventricle and ice cold
0.9% saline (200 mL for rats and 30 mL for mice) was infused into the aorta at approximately
100 mmHg pressure. The blood drained out via an incision made to the right atrium. The
brains then were quickly collected and dissected into six 2-mm-thick (1-mm-thick in mice)
coronal slices for digital imaging (Sony, Tokyo, Japan). Every 3rd slice was cut into two
halves coronally (rostral and caudal). The rostral part was embedded in Tissue-Tek (Sakura
Finetek Inc., Tokyo, Japan), snapfrozen in liquid nitrogen, and kept thereafter at 80 °C until
15-µm of sections were cut for analysis of EB-albumin extravasation (I) and 150-µm sections
were cut for mRNA analysis (IV). All the remaining slices were stained with 0.2% TTC at 37
°C (IV) and immersion-fixed in 10% formaldehyde.
3.11 ISCHEMIC LESION ASSESSMENT
3.11.1 MRI-based infarction
At acute time-points (<72 h) DWI and otherwise T2-WI were used to calculate the area or
volume of the ischemic lesion. Regions with increased signal intensity (restricted diffusion in
DWI and increased water content in T2-WI) were manually outlined. Area calculation was
based on the optic chiasmal slice, where ischemia involved both cortex and subcortex
(Figure 9). Lesion areas from all slices were summed up and multiplied by slice thickness,
yielding the uncorrected lesion volume. Afterwards, the volumetric difference between the
right and left hemisphere (due to swelling) was subtracted from the uncorrected lesion
volume, yielding the (edema) corrected lesion volume.
54
Figure 9 Ischemic lesion delineation. On the left is magnetic resonance imaging (MRI)
diffusion weighted image (DWI) taken 2 h after 90-min ischemia. On the right is digital
pictures of TTC-stained brain slices of a mouse subjected to 60-min ischemia. TTC, 2,3,5triphenyltetrazolium chloride.
3.11.2 TTC-based infarction
TTC-stained brain slices were photographed with a digital camera (Figure 9) and images
were analyzed using NIH software Image J.427 In each slice, unstained ischemic tissue,
presented as pale areas, was manually outlined. As described above, uncorrected and
edema-corrected lesion volumes were calculated.274 Lesions were reported as the
percentage of the intact hemisphere (% hemispheric lesion volume). The percentage of
edema was also reported relative to intact hemisphere.
3.12 BLOOD-BRAIN BARRIER PERMEABILITY ASSESSMENTS
3.12.1 Evans blue extravasation
Rats received intravenously (I) and mice retro-orbitally (IV) a dose of EB (3 mL/kg for rats
and 1mL/kg for mice of 2% solution; 20 mg/mL dissolved in 1% albumin, Sigma-Aldrich,
Steinheim, Germany) 20 to 25 min before collecting the brains.
Evans blue fluorescence signal intensity (I and IV) and the area of EBA extravasation (I)
were measured with a fluorescence scanner (Typhoon 9400, Amersham Biosciences,
55
Buckinghamshire, UK).428 In Study I, the area of EB extravasation was manually outlined and
the average fluorescence signal intensity within this area and of the intact hemisphere was
measured. In Study IV, because, if any, a widespread BBB leakage was expected, EB
fluorescence was calculated as the ratio of the average signal intensity of the whole brain
specimen to signal intensity of a reference point out of the hemisphere with the image
analyzer software ImageQuant (Amersham Biosciences Buckinghamshire, UK).
3.12.2 Contrast-enhanced MRI
3.12.2.1 Percentage of enhancement of the ischemic lesion
The enhancement area on postcontrast T1-weighted image represents the area of BBB
leakage. Contrast enhancement area was manually outlined (I, III) and by taking its ratio to
the lesion area and multiplying by 100 the percentage of enhanced ischemic lesion (% GdDTPA) was calculated.
3.12.2.2 Contrast-to-noise ratio of the enhancement area
Signal intensities (I) were collected from manually outlined enhancement area (Sinf), the
entire contralateral hemisphere (Snormal), and a reference point outside the brain tissue
(noise). Thereafter, signal intensity of the nonischemic hemisphere was subtracted from the
signal intensity of the enhancement area and the ratio to the noise was calculated yielding
contrast-to-noise ratio of the enhancement ((Sinf-Snormal)/noise).
3.12.2.3 Signal intensity change due to enhancement
Signal intensity values were collected (III) from the enhancement area on the postcontrast
T1-weighted image (SIpost) and from the corresponding area on the precontrast T1-weighted
image (SIpre). Signal intensity change due to Gd-DTPA enhancement was calculated by
substracting SIpre from SIpost and by taking its ratio to SIpre and multiplying by 100
((SIpost−SIpre)/SIpre×100).
3.12.2.4 The blood-to-brain transfer constant of Gd-DTPA
ROIs were manually placed (II, III) in the ischemic hemisphere on postcontrast IR-FLASH
scans in a standard manner (Figure 7). Arterial concentration of Gd-DTPA was estimated
56
from the superior sagittal sinus as described by others.396 Data collected from ROIs of each
IR-FLASH sequence were fitted to calculate T1 values and afterwards inverse T1 values were
calculated (R1). These data were further applied to Patlak plot equations396, 429 as described
above, yielding as slope the blood-to-brain transfer constant of Gd-DTPA (Ki), which is an
estimated measure of BBBP to Gd-DTPA.
3.13 QUANTITATIVE ANALYSES OF Stc1, Stc2, and Il-6 mRNA
From pooled brain slices of ischemic mice (IV), total RNA was isolated by use of TRIZOL®
Reagent (Invitrogen, Carlsbad, CA, USA). The High Capacity RNA-to-cDNA Kit (Applied
Biosystems, Foster City, CA, USA) was used to prepare cDNA and quantitative real-time
PCR was performed with the LightCycler® II instrument (Roche Diagnostics, Mannheim,
Germany) and the Maxima SYBR Green qPCR Master Mix (2X) (Fermentas, Vilnius,
Lithuania). Primers were: Stc1; 5’-ATGCTCCAAAACTCAGCAGTGATTC-3’ and 5’CAGGCTTCGGACAAGTCTGT-3’, Stc2; 5’-GCATGACGTTTCTGCACAAC and 5’CAGGTTCACAAGGTCCACAT, and Il-6 5’-CTTCCCTACTTCACAAGTCC-3’ and 5’GCCACTCCTTCTGTGACTC-3’. Stc1, Stc2, and Il-6 mRNA levels were normalized against
levels of beta-2-microglobulin, of which primers were: 5’-GCTATCCAGAAAACCCCTCA-3’
and 5’-ATGTCTCGATCCCAGTAGAC-3’. All primers were from Proligo LLC, Paris, France.
3.14 STATISTICAL ANALYSES
All parametric data are presented as mean ± SD. Normally distributed parametric data sets
were analyzed with Student’s t-test (two groups) or one-way ANOVA followed by Holm-Sidak
post hoc test (multiple groups). When normality failed, Kruskal–Wallis test followed by Dunn's
method assessed differences between multiple groups. Nonparametric data (neurological
scores) from two groups were analyzed by the Mann-Whitney U test. Repeated measures of
ANOVA followed by Holm-Sidak post hoc test examined the temporal differences of an
individual parameter. Paired t-test was used to study differences between data sets from the
same animal. MRI data of the superior sagittal sinus were fitted to an exponential curve.
Linear regression analysis of each MRI data set, which was obtained from the ROIs of T1
maps, yielded as slope the blood to brain transfer rate constant of Gd-DTPA, Ki. Spearman
correlation coefficient analysis served to identify correlations. A two-tailed value of P < 0.05
was considered significant.
57
4 RESULTS
Study I and II
These studies included a comprehensive evaluation of the BBBP following transient
occlusion of the MCA in rats. Encompassing all the hyperacute, acute, subacute, and chronic
phases of the post-reperfusion period with 15 different groups of rats (2, 4, 6, 12, 18, 24, 36,
48, and 72 h and 1, 2, 3, 4, and 5 weeks), BBBP to both large and small molecules were
quantitatively characterized, the former via the gold standard method (Evans blue
fluorescence) and the latter with gadolinium-enhanced MRI.
Animals
After the exclusions due to premature death and subarachnoid hemorrhage, 123 rats (N=6-8
per group) completed the experimental period. Control animals included eight sham-operated
rats. No significant differences arose in the physiologic parameters (mean arterial blood
pressure and temperature) between study groups.
Ischemic lesions
In all animals successful MCAO and reperfusion were ascertained with LDF, which showed a
mean CBF value 14% (±3) of the baseline during occlusion and after reperfusion 65% (±4) of
the baseline. Ischemic lesions were visualized immediately after occlusion with DWI
sequence of MRI, revealing substantial-sized cortico-subcortical lesions. Final infarct
volumes at the corresponding time-points were similar among groups (in average 0.22±0.10
cm3, P=0.42). Volumes were in good correlation with lesion areas (r=0.710, P=0.003)
calculated from the optic-chiasmal slice, which was used for BBBP quantifications. Control
animals were free of ischemic lesions.
BBB leakage to Evans blue
EB fluorescence quantification indicated that at all time-points, except for 3 and 5 weeks after
reperfusion, EB extravasated into ischemic area (P<0.001), with a slight decrease at 36 and
72 h.
58
BBB leakage to Gd-DTPA
Gd-DTPA presence in the ischemic parenchyma led to increased contrast-to-noise ratio in
the post-contrast T1-weighted images at all time-points, except for 5 weeks. This increase
was slightly lesser at the earliest time-point (25 min) of the study and at the two latest timepoints (3 and 4 weeks) of Gd-DTPA leakage.
GD-DTPA leakage estimated as blood-to-brain transfer constant (Ki) via Patlak plotting of
DCE-MRI data, indicated a sustained leakage up to 5 weeks after reperfusion (P<0.001).
Spatial pattern of BBB leakage
BBB leakage to both tracers were limited to ischemic area, but the extent of the leakage
varied depending on the time-point and the tracer, though always being smaller than the
extent of the ischemic lesion (49-90% of the ischemic lesion)(Figure 10). Starting from 72 h,
EB leaking area was smaller than Gd-DTPA leaking area (P<0.01).
Figure 10 Leaking lesion areas. The size of Evans blue (EB) and contrast agent (Gd-DTPA)
leaking areas are compared (**, P<0.01).
59
Parameters affecting BBB leakage
The severity of the ischemia extrapolated from ADC values correlated with the degree of
ischemia (r=-0.58, P=0.02), the lower the ADC value, the higher the blood-to-brain transfer
constant of Gd-DTPA. The extent of the ischemic lesion was another factor associated with
increased BBB leakage to Gd-DTPA (r=0.75, P=0.0015), larger lesions depicted a more
leaky BBB with higher Ki values. Ki values showed a trend of decrease overtime (r=-0.61,
P=0.01).
Study III
Appreciating the large standard deviations in BBBP related parameters obtained in previous
studies, this study was designed to diminish inter-animal variability and to test more
vigorously the hypothesis of continuous BBB leakage following transient ischemia. Study
included the same animal model as in the previous studies, and DCE-MRI was repeated at 5
time-points after reperfusion (2, 24, 48, and 72 h and 1 week). Signal intensity analysis and
Patlak plotting of MRI data allowed estimating BBBP to Gd-DTPA.
Ten rats with successful MCAO and reperfusion as documented by laser-Doppler
flowmetry and six sham-operated control animals were included in the study. No significant
differences emerged in physiological parameters (mean arterial blood pressure,
temperature) among animals or time-points. Baseline ischemic lesions calculated from DW
images were similar among animals (P=0.971). Uncorrected ischemic lesion volumes
increased between 2 h and 24 h and decreased thereafter, reflecting formation and
resolution of edema, respectively. A good correlation appeared between lesion volumes and
areas (r=0.767, P<0.001) calculated from the optic-chiasmal slice, which was used for BBBP
quantifications. Sham animals showed no brain pathology.
The Gd-DTPA leakage, analyzed as signal intensity change from post-contrast T1weighted images relative to precontrast images, was higher than that of shams at all timepoints (P<0.001, ANOVA), indicating a continuous leakage. Among time-points, 1 week was
associated with higher signal intensity change (P<0.001, RM-ANOVA) (Figure 11).
60
Figure 11 Monitoring of ischemic lesion and blood-brain barrier (BBB) disruption by magnetic
resonance imaging (MRI) in a representative rat. Gray scale MRI images are diffusion
images (starting from 48 h with b0, which provides T2-weighted image), colored images are
color-coded post-contrast fluid attenuated inversion recovery images. There is continuous
leakage into ischemic area, which is most pronounced at 1 week. See color scale bar for
increasing gadolinium (Gd) leakage.
A second method for estimating BBBP to Gd-DTPA used Patlak plotting of the DCE-MRI
data, which provided blood-to-brain transfer constant of Gd-DTPA, Ki. With the knowledge of
heterogeneity within the ischemic lesion, data collection applied two methods. Firstly, cortical
and subcortical parts of the ischemic lesion were analyzed as entireties, and secondly small
circular ROIs (3 per cortex and subcortex) provided the data (Figure 7B). With the first
approach, neither cortical, nor subcortical Ki values differed among time-points (P>0.05, RMANOVA), being different than those of sham animals and of contralateral values during the
whole experiment (P<0.005, ANOVA). With the second approach, a difference in Ki values
among time-points appeared only in the comparison of values from a cortical ROI (ROI-c2,
Figure 7B).
61
Study IV
This study explored the role of STC-1 in HPC and in the BBB integrity via the use of
genetically modified STC-1 deficient (STC-1-/-) mice. Transient (60 min) occlusion of the MCA
was introduced to STC-1-/- mice and wild type (WT) littermates, with or without HPC (6 h of
8% oxygenation), 24 h prior to ischemia. BBB experiments assessed EB fluorescence in
STC-1-/- mice and WT mice under normal conditions, and immediately and 24 h after hypoxia.
Real time polymerase chain reaction quantified Stc-1, Stc-2, and Il-6 mRNA with DNA
extracted from ischemic brains.
After the exclusions due to subarachnoid hemorrhage, inadequate occlusion or reperfusion,
and premature death, HPC experiments included nine to ten mice per group. 28 mice were
subjected to BBB experiments (N=4-5 per group). No differences existed in body weights or
rectal temperatures of the animals. In HPC experiments, LDF measurements ensured similar
rates of CBF reduction during MCAO (P=0.105) and of CBF recovery after reperfusion
(P=0.118).
In STC-1-/- mice and WT mice, HPC prior to ischemia resumed in equally smaller infarctions
than did ischemia only (22±10% vs. 26±8%, P=0.336). In both scenarios, STC-1-/- mice
exhibited worse neurological scores than of WT mice, although HPC improved neurological
outcome of ischemia in STC-1-/- mice (P=0.024, Figure 12).
When HPC was introduced prior to ischemia, brain mRNA expressions of Stc1 (P=0.005)
and Stc2 (P=0.035) in WT mice and of Stc2 in STC-1-/- mice (P=0.002) were increased
compared to ischemia only. After ischemia only, brain Il-6 mRNA levels differed between
STC-1-/- and WT mice (P=0.033), with 9-time lower levels in STC-1-/- mice.
EB fluorescence results were comparable between STC-1-/- and WT mice under normal
conditions (P>0.05) and the application of hypoxia did not result in increased leakage in STC/-
mice neither immediately (P>0.05), nor 24 h after hypoxia (P>0.05).
62
Figure 12 Distribution of neurological scores. STC knockout mice (STC-/-) and wild type (WT)
mice were subjected to 60-min ischemia; STC-/--HPC and WT-HPC mice received hypoxic
preconditioning (HPC) 24 h before ischemia. N=9 to 10 per group.
63
5 DISCUSSION
Stroke, the second leading cause of death worldwide,430 will affect approximately one of
every six persons,431 leaving each year 5 million people dependent on others.2 Ischemic
stroke is responsible nearly 80% of all strokes and occurs due to occlusion of a cerebral
artery. Consequently, brain region supplied by the occluded artery is left without blood flow
and undergoes complex pathophysiological events that cause irreversible damage, unless
timely reperfusion occurs. Early reperfusion (spontaneously or therapeutically) is beneficial
and limits the injury, but late reperfusion exacerbates the injury through further harmful
events. Among these, BBB disruption is the most critical.
The BBB protects central nervous system against harmful ingredients of the circulating
blood, first as a physical barrier through its structural components, second as a functional
and selective mechanical barrier through its transport mechanisms, and finally as an
enzymatic barrier. Accordingly, large molecules, including most proteins, drugs, and cellular
elements of the blood, are blocked from the central nervous system under normal conditions.
In disease conditions, such as ischemic stroke, the brain is left without this crucial guard at
varying degrees; at the extreme, massive edema and symptomatic hemorrhagic
transformation occurs.
In animal models of transient focal cerebral ischemia (ischemic stroke with reperfusion), BBB
disruption is long believed occurring in a biphasic pattern. This assumption however is based
on few studies performed more than two decades ago and are still repeatedly cited.344, 386, 400,
401
A critical analysis of these works discloses several methodological issues, which
complicates the interpretation of their data in order to achieve a common conclusion and
raises doubts on the so-called biphasic nature of the BBB leakage following ischemiareperfusion.
Knowing the time course and the degree of BBB leakage following ischemia-reperfusion is
crucially important in AIS patients for several reasons. First, patients at higher risk of
experiencing the devastating consequences of severely leaking BBB (massive edema and
64
symptomatic hemorrhagic transformation) could be detected early; second, candidate
regimens to prevent or alleviate these harmful effects could be introduced within a correct
therapeutic window, presumably early in the acute phase; third, in the later phases, when
repair mechanims take place, it enables offering the brain drugs for enhancing these
beneficial mechanisms (i.e. neurorestoration). If the BBB ceases to leak, potential drugs of
neuroprotection or neurorestoration would not reach the brain.
Biphasic BBB leakage was originally defined by Kuroiwa et al.400 as an earlier leakage
followed by a non-leaking state and a second leakage. Authors400 have applied transorbital
occlusion of the MCA to cats and they have provided detailed data on CBF values. However,
infarct sizes are missing. It is known that this ischemia model is associated with variable
outcomes.432 Only four groups of animals (for EB evaluations at 2, 3, 5 and 72 h after
reperfusion) were studied, with 6 to 11 cats per group. First group received EB injection
immediately after reperfusion and EB stayed in the circulation for two hours. In other groups,
EB was left to circulate for 30 min. BBB leakage was visually analyzed, which is the weakest
feature of the study. In all groups, except in 3 h group, EB was observed in ischemic areas.
These cats (N=11) without EB leakage however showed serum protein leakage, and this was
interpreted as the result of the first barrier opening, which occurred during zero to two hours
after reperfusion. They explained the lack of EB leakage (“refractory period”) with two
contradictory theories: Either BBB functions were fully recovered or were severely disturbed,
that EB transport through the BBB into central nervous system was inhibited.
Two following studies reporting biphasic BBB leakage, one by Belayev et al.401 and the other
by Rosenberg et al.,344 fortunately used the same stroke model (2 h ischemia with suture
MCAO) and quantitative evaluations of BBBP, but with tracers of different sizes: EB (large
tracer) and radiolabeled sucrose (small tracer), respectively. EB was left in the circulation for
one or two hours and radiolabeled sucrose for 10 minutes. Although study of Belayev et al.
did not utilize LDF-control on the ischemia, reported lesion volumes indicate that they have
induced considerable sizes of infarctions. Rosenberg et al., on the other hand, did not
provide any data on the CBF changes or infarctions induced. Time-points included in the
study of Belayev et al. generally overlap with the time-points of the study by Kuroiwa et al.400
However, at earlier time-points (0-2 h after reperfusion) when Kuroiwa et al. found the BBB
open, Belayev et al. detected no EB leakage. Starting from one to three hours after
reperfusion EB was leaking in the striatum at all the time-points evaluated, with maximum
65
leakage at 46 to 48 h after reperfusion. Rosenberg et al. studied a high number of timepoints, involving several time-points before 24 h (thus covered the long gap in the other
studies386, 400, 401) and further time-points than 72 h (which were not previously studied).
Unfortunately, the first 3 h of reperfusion, at which others were found the first BBB
opening,386, 400 was ignored. Rosenberg et al. reported two BBB openings occurring firstly at
3 h and secondly at 48 h after reperfusion. Although they did not comment on time-points
between these two extremes, they reported that BBBP was returned to normal by 14 days;
thus, one can suggest that some degree of leakage existed between 3 and 48 h postreperfusion.
Huang et al.386 induced small cortical infarcts in spontaneously hypertensive rats using
surgical distal MCAO for two hours. BBBP was evaluated with radiolabeled sucrose at timepoints mostly similar to those in the studies of Kuroiwa et al. and Belayev et al. The particular
difference was in the earliest time-point, which covered the first very minutes of reperfusion.
Huang et al. noted an increased BBBP in the neocortex at this earliest time-point that
conflicts with the study of Belayev et al. This early BBB leakage was followed by a partial
recovery at 1 and 4 h post-reperfusion, which was interpreted resulting from the closure of
the BBB. Further increases in the BBBP occurred at 22 h and at 46 h (maximal increase)
post-reperfusion.
To summarize, these four most referred studies, which suggest biphasic BBB opening after
ischemia-reperfusion, disagree on the course of biphasicity from several aspects and raise
serious concerns. Timing of the first opening is uncertain, does it occur very early (within
minutes) after reperfusion386 or within hours? 344, 401 Is there really any closure of the BBB
within leaky periods,400 does it mean a complete functional recovery in the middle of ongoing
pathologic events of ischemic cascade? Timing of presumed second opening is ambiguous
too, does it occur as early as 5 h400 or as late as 48h after reperfusion?344, 401 Does the tracer
size affect the results of post-ischemic BBBP? This last issue was not tested in above
mentioned studies.
Stimulated by these questions, the first three studies included in this thesis were performed.
A well-known transient ischemia model (suture occlusion of the MCA in rats) was assisted
with LDF and MRI to ascertain the occlusion and reperfusion, therefore to reduce outcome
variability. BBBP changes were monitored, first with a comprehensive study protocol, which
66
included 15 groups of animals, covering all the phases of post-reperfusion and avoiding long
gaps between time-points (I, II). Second, to confirm the findings of the first two studies with a
more powerful study design, BBBP changes were monitored longitudinally, in a single group
of animals (III). We quantified BBB protein permeability (via large molecule EB-albumin
extravasation) (I) and ion permeability (via small molecule, Gd-DTPA extravasation) (I; II, and
III), using the gold standard ex vivo technique (EB fluorescence) and a novel in vivo
technique (contrast-enhanced MRI), respectively.
The consistent finding in the results (I, II, and III) is that, BBB leakage following ischemiareperfusion is continuous and long-lasting, without any closure up to several weeks.
Acknowledging the complex pathophysiological changes triggered by ischemia-reperfusion,
where many mediators and mechanisms take place in a temporal manner, variations in the
degree of the BBB leakage is expected. However, the biphasic BBB leakage concept is an
oversimplification and is misleading, because it involves not only perturbations (first and
second openings) of the leakage, but also cessation of the leakage in between perturbations
(so-called closure of the BBB). Such closure was found to happen at a wide range of 0 to 12
h (Table 1), which however falls into acute phase of stroke, when ischemic injury is yet
evolving and repair mechanisms are inactive. In the studies included in this thesis (I, II, and
III) continuity of the BBB leakage was proven for both large and small molecules and with
both transversal and longitudinal study designs. Until the stage of an absolute lack of the
leakage at several weeks after reperfusion, no transient closure of the BBB (in other terms,
refractory period400) occurred. Recent MRI-based evaluations of BBBP discredit the
assumption of biphasic BBB leakage and point towards gradually increasing leakage up to
24 h (Table 2).433-439 Experimental data on long-term behaviors of the BBB following
ischemia-reperfusion are scarce and inconsistent. In rats subjected to 90-min MCAO via
suture occlusion method344 sucrose leakage (small molecule) ceased at 2 weeks. In a treevessel occlusion model in rats, after 60 min ischemia, Gd-DTPA leakage continued up to 3
weeks. Our studies (I,II) are the most comprehensive off all studies investigating BBBP after
focal brain ischemia as we covered a wide range time-points and used both a small (GdDTPA) and large molecule (EB).
In vivo BBBP imaging with CT or MRI techniques detects varying rates of increased
permeability in ischemic stroke patients (from 20 to 88%),138-140 more often in the subacute
phase. After 3 weeks parenchymal contrast enhancement tends to decrease.440 Our
67
quantitative MRI results are in agreement with human findings, that a continuous but slowly
decreasing permeability to contrast agent was detectable during 4 weeks after stroke. We
should note that our MRI experiments differ from clinical MRI practice from many aspects:
First, a high field strength MRI was used (4.7 T). Second, Gd-DTPA dose was 0.5 mmol/kg
(a considerably high dose, 5 times higher than in usual clinical practice, that even rarely
applied in experimental stroke437); and last, contrast-enhanced images were collected as late
as 30 min after Gd-DTPA bolus injection (a feature that may not be feasible in AIS patients).
All these factors improve the chance to detect even minute amounts of contrast
enhancement.
In concert with previous findings (Table 1, Figure 3), a nonsignificant decrease in BBB
leakage occurred to both large and small tracers around 24 to 36 h (I, III). Potential
explanation for this drop in the leakage is no-reflow phenomenon.441, 442 According to this
concept, plugs of neutrophils can interrupt microvascular circulation early after reperfusion,
consequently the delivery of the tracer from blood to the brain may be limited despite a
disturbed BBB. An interesting point is that at 24h post-reperfusion while the degree of
leakage decreases,443 the extent of leaky area increases.444 Another alteration observed was
at 1 week, as an increased leakage (III). However, this increase was nonsignificant, when
BBBP was evaluated encompassing the entire lesion. Only a limited cortical area was
responsible for this trend (Figure 7B, ROI-c2). Recent data concerning post-stroke
angiogenesis439, 445 suggest this finding of increased BBB leakage at 1 week post-reperfusion
a proof for regeneration, rather than being related to clinical deterioration.
Previous works disclosed that increasing the duration (therefore the severity) of the transient
ischemia results in deterioration of the BBB damage.142, 433, 434, 446 In agreement with this, our
results indicate that brain areas with lower ADC values show higher permeability and the
larger the ischemia area is, the higher the BBBP.
A growing body of data suggests increased BBBP as a predictor of hemorrhagic
transformation.140, 141, 149-151 BBBP quantification with imaging methods in stroke is yet a
developing area and awaits standardization based on large clinical trials. One unsolved issue
is the consequences of varying degrees of BBBP. In theory, if the permeability of the BBB is
not large enough for blood cellular elements to pass, hemorrhage will not occur, but BBB
leakage to much smaller molecules such as albumin, may cause edema.152 At this context,
68
BBBP evaluations with different size of tracers are of importance. In the studies included in
this thesis, post-ischemic BBB leakage to large and small tracers showed different temporal
and spatial profiles. Leakage to both tracers started early after reperfusion (at 25 min),
leakage to large molecule ceased at 3 weeks, and to small tracer at 4 weeks. When leaking
areas were compared, at 25 min post-reperfusion, surprisingly large molecule’s leakage
tended to be larger compared to small molecule’s leakage (P=0.063). Afterwards, up to 48 h,
both tracers leaked to similar extents. However, from 72 h to 2 weeks EB leaking area was
significantly smaller than contrast leaking area. Taken together, these data may suggest that
very early BBB disturbance occurring after ischemia-reperfusion is more pronounced for
larger molecules, but at the chronic phase BBB function recovers earlier for larger molecules.
Tracer size depending behavior of post-ischemic BBB was also previously noted. At 3 h after
3-hour-long transient ischemia in rats, EB extravasated into ischemic area, but red blood
cells or dextran did not.447 At 21 h, all tree elements have leaked through the BBB. In the
same animal model, at 21 h after reperfusion, small molecule leaking areas were two times
larger than large molecule leaking areas, and leakage of large molecule was associated with
higher BBBP.448 When BBB leakage to a very large molecule at 3 and 21 h post-reperfusion
was compared spatially, same authors noticed that from 3 to 21 h leaky areas were
increased by nearly 50%.449 Unfortunately, these authors focused on only 3 and 21 h postreperfusional periods. Our results on very early or late BBB behavior to different size of
tracers require confirmation with further studies.
In the studies where STC1-knockout mice were used (IV), STC1-deficiency was not
associated with increased infarct volumes, but resulted in worse neurological scores
compared to wild-type mice. Moderate-sized lesions (approximately 23% of the ipsilateral
hemisphere) caused mild neurological disability in wild-type mice, but STC1-deficient mice
with similar-sized infarcts were severely disabled or died. It is unlikely that the hemorrhagic
transformation observed in STC1-deficient mice (3 out of 10) was alone responsible for this
deterioration, because although hypoxic preconditioning was effective in reducing lesion size
in these mice, they still scored worse than wild-type mice without having hemorrhagic
transformation. We suggest that reduced expression of IL-6 (a multipotent neuroprotective
cytokine) in ischemic brains of STC1-deficient mice may have led to increased inflammation.
Elucidating the role of STC1 in preserving neurological functions after ischemia requires
further studies.
69
BBB permeability in STC1-knockout mice was disturbed neither in health, nor after hypoxia.
Although we did not examine BBBP in STC-knockout mice subjected to ischemia, we found
no increased edema formation in these mice compared to wild-type littermates. This implies
that STC1 has no major contribution in preserving BBB functions in health or disease.
70
6 SUMMARY AND CONCLUSIONS
This thesis provides strong evidence that, following transient focal ischemia, BBB leakage is
continuous (monophasic) and long-lasting. There is a transient reduction in the BBBP around
24 h post-ischemia, but it is not closed. Hence, biphasic BBB leakage for this scenario is an
over- and misused term and should be abandoned.
Findings of this thesis are clinically important from several reasons. Firstly, neurotoxicity of
thrombolysis with t-PA is increasingly appreciated and BBB protecting drugs are being
sought. BBBP imaging may be useful in detecting AIS patients that are under risk of
developing massive edema or hemorrhagic transformation. Furthermore, when testing BBB
protective strategies, quantitative BBBP evaluations are needed to monitor therapeutical
effects. Up to date, no neuroprotective or neurorestorative strategy is proven effective in AIS
patients. However, experiences from past clinical trials revealed that there is much to
improve in both preclinical studies and clinical trial designs. With these improvements
implemented, new drugs with potential long-term application may be offered to AIS patients.
From this aspect, it is crucial to elucidate when the BBB is in ischemic stroke patients and for
how long it remains so. MRI-based BBBP evaluation method used in this thesis may easily
serve for this purpose.
Stanniocalcin is yet a little-known molecule. Our findings emphasize the neuroprotective
effect of STC1, through a cross-talk with inflammatory cytokine IL-6. However, STC1 does
not appear to have a role in BBB integrity.
71
ACKNOWLEDGMENTS
This work was carried out at the Department of Neurology, Helsinki University Central
Hospital. I am sincerely thankful to all those people who have helped, guided, and supported
me during these studies.
Especially, I wish to express my gratitude to my supervisor Turgut Tatlisumak,
who gave me chance to discover the scientist and writer in me. I am deeply grateful to Turgut
for his constant help, optimism, humor, and support. He often believed in me more than I did,
he made me feel at home in a foreign country, and despite his busy schedule, he was always
available when I needed his help. I truly admire his expertise and the sharpness of his red ink
pen is outstanding! I am sincerely thankful to my other supervisor, Daniel Strbian, who
always astonished me with his limitless talents. I am deeply indebted to Daniel for not only
sharing with me his knowledge, but also for his kind friendship. Without his supportive
discussions in my miserable times, this work would not have realized.
I express my sincere gratitude to Markku Kaste and Markus Färkkilä, the
former and the present heads of the Department of Neurology and to Timo Erkinjuntti, for
the opportunity to conduct my research. I would like to thank Jari Honkaniemi and Olli
Gröhn for kindly reviewing and commenting on this thesis to improve it. Carrol Norris earns
my thanks for author editing the language for articles II and III.
I truly thank Wolf-Rudiger Schäebitz for accepting the role of being my
opponent. I believe the defense day will be memorable due to his visionary views.
I wish to express my warmest gratitude to my coworkers from Biomedicum,
Usama Abo-Ramadan, Ivan Marinkovic, Eric Pedrono, and Miia Pitkonen. It has been
pleasure to work with you and especially thanks for your friendship. Stimulating discussions
during lunch-times had impact on me! I also wish to thank coworkers, Lauri Soinne from the
Department of Neurology, Leif C. Andersson, Johan A. Westberg, and Martina
Serlachius from the Department of Pathology. Taru Puhakka earns my thanks for her skilful
technical assistance, Ertugrul Tatlisumak for co-writing, and Roger R. Reddel and Andy
C.-M. Chang for co-writing and providing knockout mice. I also deeply thank to workers of
animal facility unit of Biomedicum, especially to Anu Ahlroos. During the journey of this work
Shashank Shekhar and Pekka Hassinen have been temporary members of our lab, to both
72
I am thankful too. I warmly thank to another visiting colleague, Ufuk Emre, we all miss you in
Finland.
I owe my sincere gratitude to Sara Z. Bahar, who mentored me during my
neurology residence in Turkey and who has seen in me a potential for a scientist and
introduced me to Turgut. She and other colleagues from my former clinic, Neurology
Department of the Istanbul University, stimulated in me the enthusiasm for stroke. I also
wish to thank to the stuff of Pitäjänmäki Helthcare Centre, especially to Kristian Siekkinen
and Marjo Parkkila-Harju, the former and the present chiefs of the centre, for
encouragements and for providing me the time I needed for this work.
My deepest thanks go to my beloved family. Jarno, thanks for being always
there for me, your love and rational thinking has been the most-valuable support to me
during this work. I owe my most sincere thanks to my lovely daughter Aino, you are the joy
of my life, giving reason not to give up! Last, but not least, I thank to my parents, my brother
and my grandfather for their trust and encouragements.
This work was supported by grants from the Centre for International Mobility,
the Finnish Medical Foundation, the Oskar Öflund’s Foundation, the Orion-Farmos
Research Foundation, the Biomedicum Helsinki Foundation, and the University of
Helsinki.
73
REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
Murray CJ, Vos T, Lozano R, Naghavi M, Flaxman AD, Michaud C, Ezzati M, Shibuya K, Salomon JA,
Abdalla S, et al. Disability-adjusted life years (DALYs) for 291 diseases and injuries in 21 regions, 19902010: a systematic analysis for the Global Burden of Disease Study 2010. Lancet. 2012;380:2197-2223.
World Health Organization. Atlas of heart disease and stroke. 2004. ln: www.whi.int [Internet]. [cited
2013 Jan 27]. Available from: http://www.who.int/cardiovascular_diseases/resources/atlas/en/.
Kinsella K, He W. An aging world: 2008. 2009. ln: U.S. Census Bureau www.census.gov [Internet]. [cited
2013 Jan 27]. Available from: http://www.census.gov/prod/2009pubs/p95-09-1.pdf.
Lopez AD, Mathers CD, Ezzati M, Jamison DT, Murray CJ. Global and regional burden of disease and
risk factors, 2001: systematic analysis of population health data. Lancet. 2006;367:1747-1757.
Meretoja A, Kaste M, Roine RO, Juntunen M, Linna M, Hillbom M, Marttila R, Erila T, Rissanen A,
Sivenius J, et al. Direct costs of patients with stroke can be continuously monitored on a national level:
performance, effectiveness, and Costs of Treatment episodes in Stroke (PERFECT Stroke) Database in
Finland. Stroke. 2011;42:2007-2012.
Roger VL, Go AS, Lloyd-Jones DM, Benjamin EJ, Berry JD, Borden WB, Bravata DM, Dai S, Ford ES,
Fox CS, et al. Heart disease and stroke statistics--2012 update: a report from the American Heart
Association. Circulation. 2012;125:e2-e220.
Bamford J, Sandercock P, Dennis M, Burn J, Warlow C. Classification and natural history of clinically
identifiable subtypes of cerebral infarction. Lancet. 1991;337:1521-1526.
Adams HP, Jr., Bendixen BH, Kappelle LJ, Biller J, Love BB, Gordon DL, Marsh EE, 3rd. Classification
of subtype of acute ischemic stroke. Definitions for use in a multicenter clinical trial. TOAST. Trial of Org
10172 in Acute Stroke Treatment. Stroke. 1993;24:35-41.
Astrup J, Siesjo BK, Symon L. Thresholds in cerebral ischemia - the ischemic penumbra. Stroke.
1981;12:723-725.
Astrup J, Symon L, Branston NM, Lassen NA. Cortical evoked potential and extracellular K+ and H+ at
critical levels of brain ischemia. Stroke. 1977;8:51-57.
Heiss WD, Huber M, Fink GR, Herholz K, Pietrzyk U, Wagner R, Wienhard K. Progressive derangement
of periinfarct viable tissue in ischemic stroke. J Cereb Blood Flow Metab. 1992;12:193-203.
Hossmann KA. Viability thresholds and the penumbra of focal ischemia. Ann Neurol. 1994;36:557-565.
del Zoppo GJ, Sharp FR, Heiss WD, Albers GW. Heterogeneity in the penumbra. J Cereb Blood Flow
Metab. 2011;31:1836-1851.
Muir KW, Buchan A, von Kummer R, Rother J, Baron JC. Imaging of acute stroke. Lancet Neurol.
2006;5:755-768.
Ginsberg MD. Adventures in the pathophysiology of brain ischemia: penumbra, gene expression,
neuroprotection: the 2002 Thomas Willis Lecture. Stroke. 2003;34:214-223.
Sharp FR, Lu A, Tang Y, Millhorn DE. Multiple molecular penumbras after focal cerebral ischemia. J
Cereb Blood Flow Metab. 2000;20:1011-1032.
Heiss WD, Rosner G. Functional recovery of cortical neurons as related to degree and duration of
ischemia. Ann Neurol. 1983;14:294-301.
Huang J, Kim LJ, Poisik A, Pinsky DJ, Connolly ES, Jr. Titration of postischemic cerebral hypoperfusion
by variation of ischemic severity in a murine model of stroke. Neurosurgery. 1999;45:328-333.
Heiss WD, Kracht LW, Thiel A, Grond M, Pawlik G. Penumbral probability thresholds of cortical
flumazenil binding and blood flow predicting tissue outcome in patients with cerebral ischaemia. Brain.
2001;124:20-29.
Read SJ, Hirano T, Abbott DF, Markus R, Sachinidis JI, Tochon-Danguy HJ, Chan JG, Egan GF, Scott
AM, Bladin CF, et al. The fate of hypoxic tissue on 18F-fluoromisonidazole positron emission
tomography after ischemic stroke. Ann Neurol. 2000;48:228-235.
Read SJ, Hirano T, Abbott DF, Sachinidis JI, Tochon-Danguy HJ, Chan JG, Egan GF, Scott AM, Bladin
CF, McKay WJ, et al. Identifying hypoxic tissue after acute ischemic stroke using PET and 18Ffluoromisonidazole. Neurology. 1998;51:1617-1621.
Tagaya M, Haring HP, Stuiver I, Wagner S, Abumiya T, Lucero J, Lee P, Copeland BR, Seiffert D, del
Zoppo GJ. Rapid loss of microvascular integrin expression during focal brain ischemia reflects neuron
injury. J Cereb Blood Flow Metab. 2001;21:835-846.
Yu ZF, Bruce-Keller AJ, Goodman Y, Mattson MP. Uric acid protects neurons against excitotoxic and
74
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
metabolic insults in cell culture, and against focal ischemic brain injury in vivo. J Neurosci Res.
1998;53:613-625.
Romanos E, Planas AM, Amaro S, Chamorro A. Uric acid reduces brain damage and improves the
benefits of rt-PA in a rat model of thromboembolic stroke. J Cereb Blood Flow Metab. 2007;27:14-20.
Unal-Cevik I, Kilinc M, Can A, Gursoy-Ozdemir Y, Dalkara T. Apoptotic and necrotic death mechanisms
are concomitantly activated in the same cell after cerebral ischemia. Stroke. 2004;35:2189-2194.
Ying W, Han SK, Miller JW, Swanson RA. Acidosis potentiates oxidative neuronal death by multiple
mechanisms. J Neurochem. 1999;73:1549-1556.
Katsura K, Kristian T, Siesjo BK. Energy metabolism, ion homeostasis, and cell damage in the brain.
Biochem Soc Trans. 1994;22:991-996.
Phan TG, Wright PM, Markus R, Howells DW, Davis SM, Donnan GA. Salvaging the ischaemic
penumbra: more than just reperfusion? Clin Exp Pharmacol Physiol. 2002;29:1-10.
Dijkhuizen RM, Beekwilder JP, van der Worp HB, Berkelbach van der Sprenkel JW, Tulleken KA,
Nicolay K. Correlation between tissue depolarizations and damage in focal ischemic rat brain. Brain Res.
1999;840:194-205.
Hartings JA, Rolli ML, Lu XC, Tortella FC. Delayed secondary phase of peri-infarct depolarizations after
focal cerebral ischemia: relation to infarct growth and neuroprotection. J Neurosci. 2003;23:1160211610.
Takano K, Latour LL, Formato JE, Carano RA, Helmer KG, Hasegawa Y, Sotak CH, Fisher M. The role
of spreading depression in focal ischemia evaluated by diffusion mapping. Ann Neurol. 1996;39:308-318.
Kristian T, Siesjo BK. Calcium in ischemic cell death. Stroke. 1998;29:705-718.
Fujimura M, Morita-Fujimura Y, Murakami K, Kawase M, Chan PH. Cytosolic redistribution of
cytochrome c after transient focal cerebral ischemia in rats. J Cereb Blood Flow Metab. 1998;18:12391247.
Mergenthaler P, Dirnagl U, Meisel A. Pathophysiology of stroke: lessons from animal models. Metab
Brain Dis. 2004;19:151-167.
Love S. Apoptosis and brain ischaemia. Prog Neuropsychopharmacol Biol Psychiatry. 2003;27:267-282.
Liu T, Clark RK, McDonnell PC, Young PR, White RF, Barone FC, Feuerstein GZ. Tumor necrosis
factor-alpha expression in ischemic neurons. Stroke. 1994;25:1481-1488.
Suzuki S, Tanaka K, Nogawa S, Nagata E, Ito D, Dembo T, Fukuuchi Y. Temporal profile and cellular
localization of interleukin-6 protein after focal cerebral ischemia in rats. J Cereb Blood Flow Metab.
1999;19:1256-1262.
Schilling M, Strecker JK, Schabitz WR, Ringelstein EB, Kiefer R. Effects of monocyte chemoattractant
protein 1 on blood-borne cell recruitment after transient focal cerebral ischemia in mice. Neuroscience.
2009;161:806-812.
Touzani O, Boutin H, LeFeuvre R, Parker L, Miller A, Luheshi G, Rothwell N. Interleukin-1 influences
ischemic brain damage in the mouse independently of the interleukin-1 type I receptor. J Neurosci.
2002;22:38-43.
Schneider A, Kruger C, Steigleder T, Weber D, Pitzer C, Laage R, Aronowski J, Maurer MH, Gassler N,
Mier W, et al. The hematopoietic factor G-CSF is a neuronal ligand that counteracts programmed cell
death and drives neurogenesis. J Clin Invest. 2005;115:2083-2098.
Strecker JK, Minnerup J, Gess B, Ringelstein EB, Schabitz WR, Schilling M. Monocyte chemoattractant
protein-1-deficiency impairs the expression of IL-6, IL-1beta and G-CSF after transient focal ischemia in
mice. PLoS One. 2011;6:e25863.
Emerich DF, Dean RL, 3rd, Bartus RT. The role of leukocytes following cerebral ischemia: pathogenic
variable or bystander reaction to emerging infarct? Exp Neurol. 2002;173:168-181.
Danton GH, Dietrich WD. Inflammatory mechanisms after ischemia and stroke. J Neuropathol Exp
Neurol. 2003;62:127-136.
Heo JH, Han SW, Lee SK. Free radicals as triggers of brain edema formation after stroke. Free Radic
Biol Med. 2005;39:51-70.
Yang GY, Betz AL. Reperfusion-induced injury to the blood-brain barrier after middle cerebral artery
occlusion in rats. Stroke. 1994;25:1658-1664.
Aronowski J, Labiche LA. Perspectives on reperfusion-induced damage in rodent models of
experimental focal ischemia and role of gamma-protein kinase C. Ilar J. 2003;44:105-109.
Pan J, Konstas AA, Bateman B, Ortolano GA, Pile-Spellman J. Reperfusion injury following cerebral
ischemia: pathophysiology, MR imaging, and potential therapies. Neuroradiology. 2007;49:93-102.
Hossmann KA. Pathophysiology and therapy of experimental stroke. Cell Mol Neurobiol. 2006;26:10571083.
Schabitz WR, Steigleder T, Cooper-Kuhn CM, Schwab S, Sommer C, Schneider A, Kuhn HG.
Intravenous brain-derived neurotrophic factor enhances poststroke sensorimotor recovery and stimulates
neurogenesis. Stroke. 2007;38:2165-2172.
Popa-Wagner A, Stocker K, Balseanu AT, Rogalewski A, Diederich K, Minnerup J, Margaritescu C,
Schabitz WR. Effects of granulocyte-colony stimulating factor after stroke in aged rats. Stroke.
75
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
2010;41:1027-1031.
Beck H, Plate KH. Angiogenesis after cerebral ischemia. Acta Neuropathol. 2009;117:481-496.
Sahota P, Savitz SI. Investigational therapies for ischemic stroke: neuroprotection and neurorecovery.
Neurotherapeutics. 2011;8:434-451.
Cramer SC. Functional imaging in stroke recovery. Stroke. 2004;35:2695-2698.
The National Institute of Neurological Disorders and Stroke rt-PA Stroke Study Group. Tissue
plasminogen activator for acute ischemic stroke. N Engl J Med. 1995;333:1581-1587.
Hacke W, Kaste M, Bluhmki E, Brozman M, Davalos A, Guidetti D, Larrue V, Lees KR, Medeghri Z,
Machnig T, et al. Thrombolysis with alteplase 3 to 4.5 hours after acute ischemic stroke. N Engl J Med.
2008;359:1317-1329.
Wardlaw JM, Murray V, Berge E, del Zoppo G, Sandercock P, Lindley RL, Cohen G. Recombinant tissue
plasminogen activator for acute ischaemic stroke: an updated systematic review and meta-analysis.
Lancet. 2012;379:2364-2372.
Hacke W, Donnan G, Fieschi C, Kaste M, von Kummer R, Broderick JP, Brott T, Frankel M, Grotta JC,
Haley EC, Jr., et al. Association of outcome with early stroke treatment: pooled analysis of ATLANTIS,
ECASS, and NINDS rt-PA stroke trials. Lancet. 2004;363:768-774.
Reeves MJ, Arora S, Broderick JP, Frankel M, Heinrich JP, Hickenbottom S, Karp H, LaBresh KA,
Malarcher A, Mensah G, et al. Acute stroke care in the US: results from 4 pilot prototypes of the Paul
Coverdell National Acute Stroke Registry. Stroke. 2005;36:1232-1240.
Kaste M. Do not wait, act now. Stroke. 2007;38:3119-3120.
Demaerschalk BM, Hwang HM, Leung G. US cost burden of ischemic stroke: a systematic literature
review. Am J Manag Care. 2010;16:525-533.
Parsons M, Spratt N, Bivard A, Campbell B, Chung K, Miteff F, O'Brien B, Bladin C, McElduff P, Allen C,
et al. A randomized trial of tenecteplase versus alteplase for acute ischemic stroke. N Engl J Med.
2012;366:1099-1107.
von Kummer R, Albers GW, Mori E. The Desmoteplase in Acute Ischemic Stroke (DIAS) clinical trial
program. Int J Stroke. 2012;7:589-596.
Mauldin PD, Simpson KN, Palesch YY, Spilker JS, Hill MD, Khatri P, Broderick JP. Design of the
economic evaluation for the Interventional Management of Stroke (III) trial. Int J Stroke. 2008;3:138-144.
Mazighi M, Meseguer E, Labreuche J, Amarenco P. Bridging therapy in acute ischemic stroke: a
systematic review and meta-analysis. Stroke. 2012;43:1302-1308.
Khatri P, Hill MD, Palesch YY, Spilker J, Jauch EC, Carrozzella JA, Demchuk AM, Martin R, Mauldin P,
Dillon C, et al. Methodology of the Interventional Management of Stroke III Trial. Int J Stroke.
2008;3:130-137.
Del Zoppo GJ, Poeck K, Pessin MS, Wolpert SM, Furlan AJ, Ferbert A, Alberts MJ, Zivin JA, Wechsler L,
Busse O, et al. Recombinant tissue plasminogen activator in acute thrombotic and embolic stroke. Ann
Neurol. 1992;32:78-86.
Smith WS, Sung G, Starkman S, Saver JL, Kidwell CS, Gobin YP, Lutsep HL, Nesbit GM, Grobelny T,
Rymer MM, et al. Safety and efficacy of mechanical embolectomy in acute ischemic stroke: results of the
MERCI trial. Stroke. 2005;36:1432-1438.
Fields JD, Lutsep HL, Smith WS. Higher degrees of recanalization after mechanical thrombectomy for
acute stroke are associated with improved outcome and decreased mortality: pooled analysis of the
MERCI and Multi MERCI trials. AJNR Am J Neuroradiol. 2011;32:2170-2174.
Kidwell CS, Jahan R, Gornbein J, Alger JR, Nenov V, Ajani Z, Feng L, Meyer BC, Olson S, Schwamm
LH, et al. A Trial of Imaging Selection and Endovascular Treatment for Ischemic Stroke. N Engl J Med.
2013:[Epub ahead of print].
Penumbra Pivotal Stroke Trial Investigators. The penumbra pivotal stroke trial: safety and effectiveness
of a new generation of mechanical devices for clot removal in intracranial large vessel occlusive disease.
Stroke. 2009;40:2761-2768.
Tarr R, Hsu D, Kulcsar Z, Bonvin C, Rufenacht D, Alfke K, Stingele R, Jansen O, Frei D, Bellon R, et al.
The POST trial: initial post-market experience of the Penumbra system: revascularization of large vessel
occlusion in acute ischemic stroke in the United States and Europe. J Neurointerv Surg. 2010;2:341-344.
Penumbra Inc. Assess the Penumbra System in the Treatment of Acute Stroke (THERAPY).
In:ClinicalTrials.gov [Internet]. Bethesda (MD): National Library of Medicine (US). 2000-. [cited 2013 Jan
18]. Available from: http://clinicaltrials.gov/show/NCT01429350 NLM Identifier:NCT01429350.
Ricci S, Dinia L, Del Sette M, Anzola P, Mazzoli T, Cenciarelli S, Gandolfo C. Sonothrombolysis for
acute ischaemic stroke. Cochrane Database Syst Rev. 2012;6:CD008348.
Cerevast Therapeutics Inc. Phase 3, Randomized, Placebo-Controlled, Double-Blinded Trial of the
Combined Lysis of Thrombus With Ultrasound and Systemic Tissue Plasminogen Activator (tPA) for
Emergent Revascularization in Acute Ischemic Stroke (CLOTBUST-ER). In:ClinicalTrials.gov [Internet].
Bethesda (MD): National Library of Medicine (US). 2000-. [cited 2013 Jan 18]. Available from:
http://www.clinicaltrials.gov/show/NCT01098981 NLM Identifier: NCT01098981.
O'Collins VE, Macleod MR, Donnan GA, Horky LL, van der Worp BH, Howells DW. 1,026 experimental
76
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
102.
treatments in acute stroke. Ann Neurol. 2006;59:467-477.
Del Zoppo GJ. Why do all drugs work in animals but none in stroke patients? 1. Drugs promoting
cerebral blood flow. J Intern Med. 1995;237:79-88.
Grotta J. Why do all drugs work in animals but none in stroke patients? 2. Neuroprotective therapy. J
Intern Med. 1995;237:89-94.
Ginsberg MD. The validity of rodent brain-ischemia models is self-evident. Arch Neurol. 1996;53:10651067.
Green RA, Odergren T, Ashwood T. Animal models of stroke: do they have value for discovering
neuroprotective agents? Trends Pharmacol Sci. 2003;24:402-408.
Fisher M, Tatlisumak T. Use of animal models has not contributed to development of acute stroke
therapies: con. Stroke. 2005;36:2324-2325.
Sutherland BA, Minnerup J, Balami JS, Arba F, Buchan AM, Kleinschnitz C. Neuroprotection for
ischaemic stroke: translation from the bench to the bedside. Int J Stroke. 2012;7:407-418.
Minnerup J, Wersching H, Diederich K, Schilling M, Ringelstein EB, Wellmann J, Schabitz WR.
Methodological quality of preclinical stroke studies is not required for publication in high-impact journals.
J Cereb Blood Flow Metab. 2010;30:1619-1624.
Philip M, Benatar M, Fisher M, Savitz SI. Methodological quality of animal studies of neuroprotective
agents currently in phase II/III acute ischemic stroke trials. Stroke. 2009;40:577-581.
Minnerup J, Sutherland BA, Buchan AM, Kleinschnitz C. Neuroprotection for stroke: current status and
future perspectives. Int J Mol Sci. 2012;13:11753-11772.
Hill MD, Martin RH, Mikulis D, Wong JH, Silver FL, Terbrugge KG, Milot G, Clark WM, Macdonald RL,
Kelly ME, et al. Safety and efficacy of NA-1 in patients with iatrogenic stroke after endovascular
aneurysm repair (ENACT): a phase 2, randomised, double-blind, placebo-controlled trial. Lancet Neurol.
2012;11:942-950.
University of California San Diego. The Intravascular Cooling in the Treatment of Stroke 2/3 Trial
(ICTuS2/3). In:ClinicalTrials.gov [Internet]. Bethesda (MD): National Library of Medicine (US). 2000-.
[cited 2013 Jan 18]. Available from: http://clinicaltrials.gov/show/NCT01123161 NLM Identifier:
NCT01123161.
EuroHYP-1. 2012
Lees KR, Zivin JA, Ashwood T, Davalos A, Davis SM, Diener HC, Grotta J, Lyden P, Shuaib A,
Hardemark HG, et al. NXY-059 for acute ischemic stroke. N Engl J Med. 2006;354:588-600.
Shuaib A, Lees KR, Lyden P, Grotta J, Davalos A, Davis SM, Diener HC, Ashwood T, Wasiewski WW,
Emeribe U. NXY-059 for the treatment of acute ischemic stroke. N Engl J Med. 2007;357:562-571.
Savitz SI. A critical appraisal of the NXY-059 neuroprotection studies for acute stroke: A need for more
rigorous testing of neuroprotective agents in animal models of stroke. Exp Neurol. 2007;205:20-25.
Daiichi Pharmaceutical Co. LTD. Ebselen Trial - Pase III. In:Stroke Trials Registry. [Internet]. Dallas
(TX): The Internet Stroke Center web site. 1997-. [cited 2013 Jan 18]. Available from:
http://www.strokecenter.org/trials/clinicalstudies/ebselen-trial-phase-iii
Kikuchi K, Kawahara KI, Uchikado H, Miyagi N, Kuramoto T, Miyagi T, Morimoto Y, Ito T, Tancharoen S,
Miura N, et al. Potential of edaravone for neuroprotection in neurologic diseases that do not involve
cerebral infarction. Exp Ther Med. 2011;2:771-775.
Lampl Y, Boaz M, Gilad R, Lorberboym M, Dabby R, Rapoport A, Anca-Hershkowitz M, Sadeh M.
Minocycline treatment in acute stroke: an open-label, evaluator-blinded study. Neurology. 2007;69:14041410.
Singhealth Foundation. Neuroprotection With Minocycline Therapy for Acute Stroke Recovery Trial
(NeuMAST). In:ClinicalTrials.gov [Internet]. Bethesda (MD): National Library of Medicine (US). 2000-.
Available from: http://clinicaltrials.gov/show/NCT00930020 NLM Identifier: NCT00930020.
O'Collins VE, Macleod MR, Donnan GA, Howells DW. Evaluation of combination therapy in animal
models of cerebral ischemia. J Cereb Blood Flow Metab. 2012;32:585-597.
Ehrenreich H, Weissenborn K, Prange H, Schneider D, Weimar C, Wartenberg K, Schellinger PD, Bohn
M, Becker H, Wegrzyn M, et al. Recombinant human erythropoietin in the treatment of acute ischemic
stroke. Stroke. 2009;40:e647-656.
Amaro S, Canovas D, Castellanos M, Gallego J, Marti-Febregas J, Segura T, Chamorro A. The URICOICTUS study, a phase 3 study of combined treatment with uric acid and rtPA administered intravenously
in acute ischaemic stroke patients within the first 4.5 h of onset of symptoms. Int J Stroke. 2010;5:325328.
The Intravascular Cooling in the treatment of stroke 2/3 trial (ictus2/3). 2010
Cook DJ, Teves L, Tymianski M. Treatment of stroke with a PSD-95 inhibitor in the gyrencephalic
primate brain. Nature. 2012;483:213-217.
Savitz SI, Schabitz WR. Reviving neuroprotection using a new approach: targeting postsynaptic density95 to arrest glutamate excitotoxicity. Stroke. 2012;43:3411-3412.
Lo EH. A new penumbra: transitioning from injury into repair after stroke. Nat Med. 2008;14:497-500.
Zhao BQ, Tejima E, Lo EH. Neurovascular proteases in brain injury, hemorrhage and remodeling after
77
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.
stroke. Stroke. 2007;38:748-752.
Savitz SI, Schabitz WR. A Critique of SAINT II: wishful thinking, dashed hopes, and the future of
neuroprotection for acute stroke. Stroke. 2008;39:1389-1391.
Minnerup J, Sevimli S, Schabitz WR. Granulocyte-colony stimulating factor for stroke treatment:
mechanisms of action and efficacy in preclinical studies. Exp Transl Stroke Med. 2009;1:2.
Minnerup J, Heidrich J, Wellmann J, Rogalewski A, Schneider A, Schabitz WR. Meta-analysis of the
efficacy of granulocyte-colony stimulating factor in animal models of focal cerebral ischemia. Stroke.
2008;39:1855-1861.
Floel A, Warnecke T, Duning T, Lating Y, Uhlenbrock J, Schneider A, Vogt G, Laage R, Koch W, Knecht
S, et al. Granulocyte-colony stimulating factor (G-CSF) in stroke patients with concomitant vascular
disease--a randomized controlled trial. PLoS One. 2011;6:e19767.
Lang W, Stadler CH, Poljakovic Z, Fleet D. A prospective, randomized, placebo-controlled, double-blind
trial about safety and efficacy of combined treatment with alteplase (rt-PA) and Cerebrolysin in acute
ischaemic hemispheric stroke. Int J Stroke. 2012
Hjort N, Butcher K, Davis SM, Kidwell CS, Koroshetz WJ, Rother J, Schellinger PD, Warach S,
Ostergaard L. Magnetic resonance imaging criteria for thrombolysis in acute cerebral infarct. Stroke.
2005;36:388-397.
Urbach H, Flacke S, Keller E, Textor J, Berlis A, Hartmann A, Reul J, Solymosi L, Schild HH.
Detectability and detection rate of acute cerebral hemisphere infarcts on CT and diffusion-weighted MRI.
Neuroradiology. 2000;42:722-727.
Mullins ME, Schaefer PW, Sorensen AG, Halpern EF, Ay H, He J, Koroshetz WJ, Gonzalez RG. CT and
conventional and diffusion-weighted MR imaging in acute stroke: study in 691 patients at presentation to
the emergency department. Radiology. 2002;224:353-360.
Fiebach J, Jansen O, Schellinger P, Knauth M, Hartmann M, Heiland S, Ryssel H, Pohlers O, Hacke W,
Sartor K. Comparison of CT with diffusion-weighted MRI in patients with hyperacute stroke.
Neuroradiology. 2001;43:628-632.
Fiebach JB, Schellinger PD, Jansen O, Meyer M, Wilde P, Bender J, Schramm P, Juttler E, Oehler J,
Hartmann M, et al. CT and diffusion-weighted MR imaging in randomized order: diffusion-weighted
imaging results in higher accuracy and lower interrater variability in the diagnosis of hyperacute ischemic
stroke. Stroke. 2002;33:2206-2210.
Barber PA, Darby DG, Desmond PM, Gerraty RP, Yang Q, Li T, Jolley D, Donnan GA, Tress BM, Davis
SM. Identification of major ischemic change. Diffusion-weighted imaging versus computed tomography.
Stroke. 1999;30:2059-2065.
Lansberg MG, Albers GW, Beaulieu C, Marks MP. Comparison of diffusion-weighted MRI and CT in
acute stroke. Neurology. 2000;54:1557-1561.
Ay H, Oliveira-Filho J, Buonanno FS, Ezzeddine M, Schaefer PW, Rordorf G, Schwamm LH, Gonzalez
RG, Koroshetz WJ. Diffusion-weighted imaging identifies a subset of lacunar infarction associated with
embolic source. Stroke. 1999;30:2644-2650.
Wessels T, Rottger C, Jauss M, Kaps M, Traupe H, Stolz E. Identification of embolic stroke patterns by
diffusion-weighted MRI in clinically defined lacunar stroke syndromes. Stroke. 2005;36:757-761.
Baird AE, Lovblad KO, Schlaug G, Edelman RR, Warach S. Multiple acute stroke syndrome: marker of
embolic disease? Neurology. 2000;54:674-678.
Warach S, Gaa J, Siewert B, Wielopolski P, Edelman RR. Acute human stroke studied by whole brain
echo planar diffusion-weighted magnetic resonance imaging. Ann Neurol. 1995;37:231-241.
Schlaug G, Benfield A, Baird AE, Siewert B, Lovblad KO, Parker RA, Edelman RR, Warach S. The
ischemic penumbra: operationally defined by diffusion and perfusion MRI. Neurology. 1999;53:15281537.
Davis SM, Donnan GA, Parsons MW, Levi C, Butcher KS, Peeters A, Barber PA, Bladin C, De Silva DA,
Byrnes G, et al. Effects of alteplase beyond 3 h after stroke in the Echoplanar Imaging Thrombolytic
Evaluation Trial (EPITHET): a placebo-controlled randomised trial. Lancet Neurol. 2008;7:299-309.
Schabitz WR. MR mismatch is useful for patient selection for thrombolysis: no. Stroke. 2009;40:29082909.
Nagakane Y, Christensen S, Brekenfeld C, Ma H, Churilov L, Parsons MW, Levi CR, Butcher KS,
Peeters A, Barber PA, et al. EPITHET: Positive Result After Reanalysis Using Baseline DiffusionWeighted Imaging/Perfusion-Weighted Imaging Co-Registration. Stroke. 2011;42:59-64.
Fink JN, Kumar S, Horkan C, Linfante I, Selim MH, Caplan LR, Schlaug G. The stroke patient who woke
up: clinical and radiological features, including diffusion and perfusion MRI. Stroke. 2002;33:988-993.
Lago A, Geffner D, Tembl J, Landete L, Valero C, Baquero M. Circadian variation in acute ischemic
stroke: a hospital-based study. Stroke. 1998;29:1873-1875.
Thomalla G, Cheng B, Ebinger M, Hao Q, Tourdias T, Wu O, Kim JS, Breuer L, Singer OC, Warach S, et
al. DWI-FLAIR mismatch for the identification of patients with acute ischaemic stroke within 4.5 h of
symptom onset (PRE-FLAIR): a multicentre observational study. Lancet Neurol. 2011;10:978-986.
Sorensen AG, Copen WA, Ostergaard L, Buonanno FS, Gonzalez RG, Rordorf G, Rosen BR, Schwamm
78
127.
128.
129.
130.
131.
132.
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
150.
LH, Weisskoff RM, Koroshetz WJ. Hyperacute stroke: simultaneous measurement of relative cerebral
blood volume, relative cerebral blood flow, and mean tissue transit time. Radiology. 1999;210:519-527.
Neumann-Haefelin T, Wittsack HJ, Wenserski F, Siebler M, Seitz RJ, Modder U, Freund HJ. Diffusionand perfusion-weighted MRI. The DWI/PWI mismatch region in acute stroke. Stroke. 1999;30:15911597.
Grandin CB, Duprez TP, Smith AM, Oppenheim C, Peeters A, Robert AR, Cosnard G. Which MRderived perfusion parameters are the best predictors of infarct growth in hyperacute stroke?
Comparative study between relative and quantitative measurements. Radiology. 2002;223:361-370.
Kane I, Carpenter T, Chappell F, Rivers C, Armitage P, Sandercock P, Wardlaw J. Comparison of 10
different magnetic resonance perfusion imaging processing methods in acute ischemic stroke: effect on
lesion size, proportion of patients with diffusion/perfusion mismatch, clinical scores, and radiologic
outcomes. Stroke. 2007;38:3158-3164.
Galinovic I, Brunecker P, Ostwaldt AC, Soemmer C, Hotter B, Fiebach JB. Fully automated
postprocessing carries a risk of substantial overestimation of perfusion deficits in acute stroke magnetic
resonance imaging. Cerebrovasc Dis. 2011;31:408-413.
Kranz PG, Eastwood JD. Does diffusion-weighted imaging represent the ischemic core? An evidencebased systematic review. AJNR Am J Neuroradiol. 2009;30:1206-1212.
Maulaz A, Piechowski-Jozwiak B, Michel P, Bogousslavsky J. Selecting patients for early stroke
treatment with penumbra images. Cerebrovasc Dis. 2005;20 Suppl 2:19-24.
Campbell BC, Purushotham A, Christensen S, Desmond PM, Nagakane Y, Parsons MW, Lansberg MG,
Mlynash M, Straka M, De Silva DA, et al. The infarct core is well represented by the acute diffusion
lesion: sustained reversal is infrequent. J Cereb Blood Flow Metab. 2012;32:50-56.
Wing SD, Norman D, Pollock JA, Newton TH. Contrast enhancement of cerebral infarcts in computed
tomography. Radiology. 1976;121:89-92.
Lee KF, Chambers RA, Diamond C, Park CH, Thompson NL, Jr., Schnapf D, Pripstein S. Evaluation of
cerebral infarction by computed tomography with special emphasis on microinfarction. Neuroradiology.
1978;16:156-158.
Norton GA, Kishore PR, Lin J. CT contrast enhancement in cerebral infarction. AJR Am J Roentgenol.
1978;131:881-885.
Hornig CR, Busse O, Buettner T, Dorndorf W, Agnoli A, Akengin Z. CT contrast enhancement on brain
scans and blood-CSF barrier disturbances in cerebral ischemic infarction. Stroke. 1985;16:268-273.
Bang OY, Saver JL, Alger JR, Shah SH, Buck BH, Starkman S, Ovbiagele B, Liebeskind DS. Patterns
and predictors of blood-brain barrier permeability derangements in acute ischemic stroke. Stroke.
2009;40:454-461.
Kassner A, Roberts T, Taylor K, Silver F, Mikulis D. Prediction of hemorrhage in acute ischemic stroke
using permeability MR imaging. AJNR Am J Neuroradiol. 2005;26:2213-2217.
Lin K, Kazmi KS, Law M, Babb J, Peccerelli N, Pramanik BK. Measuring elevated microvascular
permeability and predicting hemorrhagic transformation in acute ischemic stroke using first-pass
dynamic perfusion CT imaging. AJNR Am J Neuroradiol. 2007;28:1292-1298.
Hayman LA, Evans RA, Bastion FO, Hinck VC. Delayed high dose contrast CT: identifying patients at
risk of massive hemorrhagic infarction. AJR Am J Roentgenol. 1981;136:1151-1159.
Knight RA, Barker PB, Fagan SC, Li Y, Jacobs MA, Welch KM. Prediction of impending hemorrhagic
transformation in ischemic stroke using magnetic resonance imaging in rats. Stroke. 1998;29:144-151.
Dijkhuizen RM, Asahi M, Wu O, Rosen BR, Lo EH. Delayed rt-PA treatment in a rat embolic stroke
model: diagnosis and prognosis of ischemic injury and hemorrhagic transformation with magnetic
resonance imaging. J Cereb Blood Flow Metab. 2001;21:964-971.
Neumann-Haefelin C, Brinker G, Uhlenkuken U, Pillekamp F, Hossmann KA, Hoehn M. Prediction of
hemorrhagic transformation after thrombolytic therapy of clot embolism: an MRI investigation in rat brain.
Stroke. 2002;33:1392-1398.
Jiang Q, Zhang RL, Zhang ZG, Knight RA, Ewing JR, Ding G, Lu M, Arniego P, Zhang L, Hu J, et al.
Magnetic resonance imaging characterization of hemorrhagic transformation of embolic stroke in the rat.
J Cereb Blood Flow Metab. 2002;22:559-568.
Vo KD, Santiago F, Lin W, Hsu CY, Lee Y, Lee JM. MR imaging enhancement patterns as predictors of
hemorrhagic transformation in acute ischemic stroke. AJNR Am J Neuroradiol. 2003;24:674-679.
Kim EY, Na DG, Kim SS, Lee KH, Ryoo JW, Kim HK. Prediction of hemorrhagic transformation in acute
ischemic stroke: role of diffusion-weighted imaging and early parenchymal enhancement. AJNR Am J
Neuroradiol. 2005;26:1050-1055.
Latour LL, Kang DW, Ezzeddine MA, Chalela JA, Warach S. Early blood-brain barrier disruption in
human focal brain ischemia. Ann Neurol. 2004;56:468-477.
Aviv RI, d'Esterre CD, Murphy BD, Hopyan JJ, Buck B, Mallia G, Li V, Zhang L, Symons SP, Lee TY.
Hemorrhagic transformation of ischemic stroke: prediction with CT perfusion. Radiology. 2009;250:867877.
Hom J, Dankbaar JW, Soares BP, Schneider T, Cheng SC, Bredno J, Lau BC, Smith W, Dillon WP,
79
151.
152.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
165.
166.
167.
168.
169.
170.
171.
172.
173.
174.
Wintermark M. Blood-brain barrier permeability assessed by perfusion CT predicts symptomatic
hemorrhagic transformation and malignant edema in acute ischemic stroke. AJNR Am J Neuroradiol.
2011;32:41-48.
Hom J, Dankbaar JW, Schneider T, Cheng SC, Bredno J, Wintermark M. Optimal duration of acquisition
for dynamic perfusion CT assessment of blood-brain barrier permeability using the Patlak model. AJNR
Am J Neuroradiol. 2009;30:1366-1370.
Bektas H, Wu TC, Kasam M, Harun N, Sitton CW, Grotta JC, Savitz SI. Increased blood-brain barrier
permeability on perfusion CT might predict malignant middle cerebral artery infarction. Stroke.
2010;41:2539-2544.
Henning EC, Latour LL, Warach S. Verification of enhancement of the CSF space, not parenchyma, in
acute stroke patients with early blood-brain barrier disruption. J Cereb Blood Flow Metab. 2008;28:882886.
Kohrmann M, Struffert T, Frenzel T, Schwab S, Doerfler A. The hyperintense acute reperfusion marker
on fluid-attenuated inversion recovery magnetic resonance imaging is caused by gadolinium in the
cerebrospinal fluid. Stroke. 2012;43:259-261.
Warach S, Latour LL. Evidence of reperfusion injury, exacerbated by thrombolytic therapy, in human
focal brain ischemia using a novel imaging marker of early blood-brain barrier disruption. Stroke.
2004;35:2659-2661.
Kidwell CS, Latour L, Saver JL, Alger JR, Starkman S, Duckwiler G, Jahan R, Vinuela F, Kang DW,
Warach S. Thrombolytic toxicity: blood brain barrier disruption in human ischemic stroke. Cerebrovasc
Dis. 2008;25:338-343.
Rozanski M, Ebinger M, Schmidt WU, Hotter B, Pittl S, Heuschmann PU, Jungehuelsing JG, Fiebach JB.
Hyperintense acute reperfusion marker on FLAIR is not associated with early haemorrhagic
transformation in the elderly. Eur Radiol. 2010;20:2990-2996.
Bang OY, Buck BH, Saver JL, Alger JR, Yoon SR, Starkman S, Ovbiagele B, Kim D, Ali LK, Sanossian
N, et al. Prediction of hemorrhagic transformation after recanalization therapy using T2*-permeability
magnetic resonance imaging. Ann Neurol. 2007;62:170-176.
Wu S, Thornhill RE, Chen S, Rammo W, Mikulis DJ, Kassner A. Relative recirculation: a fast, model-free
surrogate for the measurement of blood-brain barrier permeability and the prediction of hemorrhagic
transformation in acute ischemic stroke. Invest Radiol. 2009;44:662-668.
Thornhill RE, Chen S, Rammo W, Mikulis DJ, Kassner A. Contrast-enhanced MR imaging in acute
ischemic stroke: T2* measures of blood-brain barrier permeability and their relationship to T1 estimates
and hemorrhagic transformation. AJNR Am J Neuroradiol. 2010;31:1015-1022.
Hossmann KA. Animal models of cerebral ischemia. 1. Review of literature. Cerebrovasc Dis. 1991;1
(suppl 1):2-15.
Sevick RJ, Kucharczyk J, Mintorovitch J, Moseley ME, Derugin N, Norman D. Diffusion-weighted MR
imaging and T2-weighted MR imaging in acute cerebral ischaemia: comparison and correlation with
histopathology. Acta Neurochir Suppl (Wien). 1990;51:210-212.
Fisher M, Feuerstein G, Howells DW, Hurn PD, Kent TA, Savitz SI, Lo EH. Update of the stroke therapy
academic industry roundtable preclinical recommendations. Stroke. 2009;40:2244-2250.
Stroke Therapy Academic Industry Roundtable. Recommendations for standards regarding preclinical
neuroprotective and restorative drug development. Stroke. 1999;30:2752-2758.
Macrae IM. Preclinical stroke research--advantages and disadvantages of the most common rodent
models of focal ischaemia. Br J Pharmacol. 2011;164:1062-1078.
Howells DW, Porritt MJ, Rewell SS, O'Collins V, Sena ES, van der Worp HB, Traystman RJ, Macleod
MR. Different strokes for different folks: the rich diversity of animal models of focal cerebral ischemia. J
Cereb Blood Flow Metab. 2010;30:1412-1431.
Sicard KM, Fisher M. Animal models of focal brain ischemia. Exp Transl Stroke Med. 2009;1:7.
Durukan A, Tatlisumak T. Acute ischemic stroke: overview of major experimental rodent models,
pathophysiology, and therapy of focal cerebral ischemia. Pharmacol Biochem Behav. 2007;87:179-197.
Tatlisumak T, Li F, Fisher M. Animal models of ischemic stroke. In: Bhardwaj A, Alkayed NJ, Kirsch JR,
Traystman RJ, eds. Acute stroke, bench to bedside. New York: Informa Healthcare; 2006:171-186.
Kassem-Moussa H, Graffagnino C. Nonocclusion and spontaneous recanalization rates in acute
ischemic stroke: a review of cerebral angiography studies. Arch Neurol. 2002;59:1870-1873.
Rha JH, Saver JL. The impact of recanalization on ischemic stroke outcome: a meta-analysis. Stroke.
2007;38:967-973.
Grotta J. Stroke treatment in the human versus animal models. In: Ginsberg MD, Bogousslavsky J, eds.
Cerebrovascular disease: pathophysiology, diagnosis, and management. Blackwell Scientific
Publications; 1998.
Hill NC, Millikan CH, Wakim KG, Sayre GP. Studies in cerebrovascular disease. VII. Experimental
production of cerebral infarction by intracarotid injection of homologous blood clot; preliminary report.
Mayo Clin Proc. 1955;30:625-633.
Kudo M, Aoyama A, Ichimori S, Fukunaga N. An animal model of cerebral infarction. Homologous blood
80
175.
176.
177.
178.
179.
180.
181.
182.
183.
184.
185.
186.
187.
188.
189.
190.
191.
192.
193.
194.
195.
196.
197.
198.
199.
clot emboli in rats. Stroke. 1982;13:505-508.
Kaneko D, Nakamura N, Ogawa T. Cerebral infarction in rats using homologous blood emboli:
development of a new experimental model. Stroke. 1985;16:76-84.
Overgaard K, Sereghy T, Boysen G, Pedersen H, Hoyer S, Diemer NH. A rat model of reproducible
cerebral infarction using thrombotic blood clot emboli. J Cereb Blood Flow Metab. 1992;12:484-490.
Takano K, Carano RA, Tatlisumak T, Meiler M, Sotak CH, Kleinert HD, Fisher M. Efficacy of intra-arterial
and intravenous prourokinase in an embolic stroke model evaluated by diffusion-perfusion magnetic
resonance imaging. Neurology. 1998;50:870-875.
Busch E, Kruger K, Hossmann KA. Improved model of thromboembolic stroke and rt-PA induced
reperfusion in the rat. Brain Res. 1997;778:16-24.
Zhang Z, Chopp M, Zhang RL, Goussev A. A mouse model of embolic focal cerebral ischemia. J Cereb
Blood Flow Metab. 1997;17:1081-1088.
Wang CX, Yang Y, Yang T, Shuaib A. A focal embolic model of cerebral ischemia in rats: introduction
and evaluation. Brain Res Brain Res Protoc. 2001;7:115-120.
Orset C, Macrez R, Young AR, Panthou D, Angles-Cano E, Maubert E, Agin V, Vivien D. Mouse model
of in situ thromboembolic stroke and reperfusion. Stroke. 2007;38:2771-2778.
Ren M, Lin ZJ, Qian H, Choudhury GR, Liu R, Liu H, Yang SH. Embolic middle cerebral artery occlusion
model using thrombin and fibrinogen composed clots in rat. J Neurosci Methods. 2012;211:296-304.
Overgaard K. Thrombolytic therapy in experimental embolic stroke. Cerebrovasc Brain Metab Rev.
1994;6:257-286.
Brinker G, Franke C, Hoehn M, Uhlenkuken U, Hossmann KA. Thrombolysis of cerebral clot embolism in
rat: effect of treatment delay. Neuroreport. 1999;10:3269-3272.
Zhang L, Zhang ZG, Zhang C, Zhang RL, Chopp M. Intravenous administration of a GPIIb/IIIa receptor
antagonist extends the therapeutic window of intra-arterial tenecteplase-tissue plasminogen activator in
a rat stroke model. Stroke. 2004;35:2890-2895.
Amaro S, Chamorro A. Translational stroke research of the combination of thrombolysis and antioxidant
therapy. Stroke. 2011;42:1495-1499.
Culp WC, Woods SD, Brown AT, Lowery JD, Hennings LJ, Skinner RD, Borrelli MJ, Roberson PK. Three
variations in rabbit angiographic stroke models. J Neurosci Methods. 2012;212:322-328.
Gounis MJ, Nogueira RG, Mehra M, Chueh J, Wakhloo AK. A thromboembolic model for the efficacy and
safety evaluation of combined mechanical and pharmacologic revascularization strategies. J Neurointerv
Surg. 2012
Koizumi J, Yoshida Y, Nakazawa T, Ooneda G. Experimental studies of ischemic brain edema: 1. A new
experimental model of cerebral embolism in rats in which recirculation can be introduced in the ischemic
area. Jpn J Stroke. 1986;8:1-8.
Takano K, Tatlisumak T, Bergmann AG, Gibson D. G. 3rd, Fisher M. Reproducibility and reliability of
middle cerebral artery occlusion using a silicone-coated suture (Koizumi) in rats. J Neurol Sci.
1997;153:8-11.
Li F, Han S, Tatlisumak T, Carano RA, Irie K, Sotak CH, Fisher M. A new method to improve in-bore
middle cerebral artery occlusion in rats: demonstration with diffusion- and perfusion-weighted imaging.
Stroke. 1998;29:1715-1719.
Longa EZ, Weinstein PR, Carlson S, Cummins R. Reversible middle cerebral artery occlusion without
craniectomy in rats. Stroke. 1989;20:84-91.
Belayev L, Alonso OF, Busto R, Zhao W, Ginsberg MD. Middle cerebral artery occlusion in the rat by
intraluminal suture. Neurological and pathological evaluation of an improved model. Stroke.
1996;27:1616-1622.
Abraham H, Somogyvari-Vigh A, Maderdrut JL, Vigh S, Arimura A. Filament size influences temperature
changes and brain damage following middle cerebral artery occlusion in rats. Exp Brain Res.
2002;142:131-138.
Hata R, Mies G, Wiessner C, Fritze K, Hesselbarth D, Brinker G, Hossmann KA. A reproducible model of
middle cerebral artery occlusion in mice: hemodynamic, biochemical, and magnetic resonance imaging.
J Cereb Blood Flow Metab. 1998;18:367-375.
Laing RJ, Jakubowski J, Laing RW. Middle cerebral artery occlusion without craniectomy in rats. Which
method works best? Stroke. 1993;24:294-297.
Shimamura N, Matchett G, Tsubokawa T, Ohkuma H, Zhang J. Comparison of silicon-coated nylon
suture to plain nylon suture in the rat middle cerebral artery occlusion model. J Neurosci Methods.
2006;156:161-165.
Zarow GJ, Karibe H, States BA, Graham SH, Weinstein PR. Endovascular suture occlusion of the middle
cerebral artery in rats: effect of suture insertion distance on cerebral blood flow, infarct distribution and
infarct volume. Neurol Res. 1997;19:409-416.
He Z, Yamawaki T, Yang S, Day AL, Simpkins JW, Naritomi H. Experimental model of small deep
infarcts involving the hypothalamus in rats: changes in body temperature and postural reflex. Stroke.
1999;30:2743-2751.
81
200.
201.
202.
203.
204.
205.
206.
207.
208.
209.
210.
211.
212.
213.
214.
215.
216.
217.
218.
219.
220.
221.
222.
223.
224.
225.
226.
Schmid-Elsaesser R, Zausinger S, Hungerhuber E, Baethmann A, Reulen HJ. A critical reevaluation of
the intraluminal thread model of focal cerebral ischemia: evidence of inadvertent premature reperfusion
and subarachnoid hemorrhage in rats by laser-Doppler flowmetry. Stroke. 1998;29:2162-2170.
Li F, Omae T, Fisher M. Spontaneous hyperthermia and its mechanism in the intraluminal suture middle
cerebral artery occlusion model of rats. Stroke. 1999;30:2464-2470.
Meng X, Fisher M, Shen Q, Sotak CH, Duong TQ. Characterizing the diffusion/perfusion mismatch in
experimental focal cerebral ischemia. Ann Neurol. 2004;55:207-212.
Henninger N, Sicard KM, Schmidt KF, Bardutzky J, Fisher M. Comparison of ischemic lesion evolution in
embolic versus mechanical middle cerebral artery occlusion in Sprague Dawley rats using diffusion and
perfusion imaging. Stroke. 2006;37:1283-1287.
Aoki T, Sumii T, Mori T, Wang X, Lo EH. Blood-brain barrier disruption and matrix metalloproteinase-9
expression during reperfusion injury: mechanical versus embolic focal ischemia in spontaneously
hypertensive rats. Stroke. 2002;33:2711-2717.
Hossmann KA. The two pathophysiologies of focal brain ischemia: implications for translational stroke
research. J Cereb Blood Flow Metab. 2012;32:1310-1316.
Pulsinelli W, Jacewicz M. Animal models of brain ischemia. In: Barnett HJ, Mohr JP, Stein BM, Yatsu
FM, eds. Stroke: Pathophysiology, diagnosis, and management. New York, NY, USA: Churchill
Livingstone; 1992:49-67.
Buchan AM, Xue D, Slivka A. A new model of temporary focal neocortical ischemia in the rat. Stroke.
1992;23:273-279.
Shigeno T, Teasdale GM, McCulloch J, Graham DI. Recirculation model following MCA occlusion in rats.
Cerebral blood flow, cerebrovascular permeability, and brain edema. J Neurosurg. 1985;63:272-277.
Robinson RG, Shoemaker WJ, Schlumpf M, Valk T, Bloom FE. Effect of experimental cerebral infarction
in rat brain on catecholamines and behaviour. Nature. 1975;255:332-334.
Tamura A, Graham DI, McCulloch J, Teasdale GM. Focal cerebral ischaemia in the rat: 1. Description of
technique and early neuropathological consequences following middle cerebral artery occlusion. J Cereb
Blood Flow Metab. 1981;1:53-60.
Bederson JB, Pitts LH, Tsuji M, Nishimura MC, Davis RL, Bartkowski H. Rat middle cerebral artery
occlusion: evaluation of the model and development of a neurologic examination. Stroke. 1986;17:472476.
Brint S, Jacewicz M, Kiessling M, Tanabe J, Pulsinelli W. Focal brain ischemia in the rat: methods for
reproducible neocortical infarction using tandem occlusion of the distal middle cerebral and ipsilateral
common carotid arteries. J Cereb Blood Flow Metab. 1988;8:474-485.
Liu TH, Beckman JS, Freeman BA, Hogan EL, Hsu CY. Polyethylene glycol-conjugated superoxide
dismutase and catalase reduce ischemic brain injury. Am J Physiol. 1989;256:H589-593.
Yanamoto H, Nagata I, Niitsu Y, Xue JH, Zhang Z, Kikuchi H. Evaluation of MCAO stroke models in
normotensive rats: standardized neocortical infarction by the 3VO technique. Exp Neurol. 2003;182:261274.
Lauer KK, Shen H, Stein EA, Ho KC, Kampine JP, Hudetz AG. Focal cerebral ischemia in rats produced
by intracarotid embolization with viscous silicone. Neurol Res. 2002;24:181-190.
Purdy PD, Devous MD, Sr., Batjer HH, White CL, 3rd, Meyer Y, Samson DS. Microfibrillar collagen
model of canine cerebral infarction. Stroke. 1989;20:1361-1367.
Yang Y, Yang T, Li Q, Wang CX, Shuaib A. A new reproducible focal cerebral ischemia model by
introduction of polyvinylsiloxane into the middle cerebral artery: a comparison study. J Neurosci
Methods. 2002;118:199-206.
Molnar L, Hegedus K, Fekete I. A new model for inducing transient cerebral ischemia and subsequent
reperfusion in rabbits without craniectomy. Stroke. 1988;19:1262-1266.
Rapp JH, Pan XM, Yu B, Swanson RA, Higashida RT, Simpson P, Saloner D. Cerebral ischemia and
infarction from atheroemboli <100 microm in size. Stroke. 2003;34:1976-1980.
Roos MW, Ericsson A, Berg M, Sperber GO, Sjoquist M, Meyerson BJ. Functional evaluation of cerebral
microembolization in the rat. Brain Res. 2003;961:15-21.
Zivin JA, DeGirolami U, Kochhar A, Lyden PD, Mazzarella V, Hemenway CC, Henry ME. A model for
quantitative evaluation of embolic stroke therapy. Brain Res. 1987;435:305-309.
Fukuchi K, Kusuoka H, Watanabe Y, Nishimura T. Correlation of sequential MR images of microsphereinduced cerebral ischemia with histologic changes in rats. Invest Radiol. 1999;34:698-703.
Hennings LJ, Flores R, Roberson PK, Brown A, Lowery J, Borrelli M, Culp WC. Persistent penumbra in a
rabbit stroke model: incidence and histologic characteristics. Stroke Res Treat. 2011;2011:764830.
Mayzel-Oreg O, Omae T, Kazemi M, Li F, Fisher M, Cohen Y, Sotak CH. Microsphere-induced embolic
stroke: an MRI study. Magn Reson Med. 2004;51:1232-1238.
Watson BD, Dietrich WD, Busto R, Wachtel MS, Ginsberg MD. Induction of reproducible brain infarction
by photochemically initiated thrombosis. Ann Neurol. 1985;17:497-504.
Dietrich WD, Watson BD, Busto R, Ginsberg MD, Bethea JR. Photochemically induced cerebral
infarction. I. Early microvascular alterations. Acta Neuropathol (Berl). 1987;72:315-325.
82
227.
228.
229.
230.
231.
232.
233.
234.
235.
236.
237.
238.
239.
240.
241.
242.
243.
244.
245.
246.
247.
248.
249.
250.
251.
252.
Dietrich WD, Busto R, Watson BD, Scheinberg P, Ginsberg MD. Photochemically induced cerebral
infarction. II. Edema and blood-brain barrier disruption. Acta Neuropathol (Berl). 1987;72:326-334.
Wester P, Watson BD, Prado R, Dietrich WD. A photothrombotic 'ring' model of rat stroke-in-evolution
displaying putative penumbral inversion. Stroke. 1995;26:444-450.
Hilger T, Blunk JA, Hoehn M, Mies G, Wester P. Characterization of a novel chronic photothrombotic ring
stroke model in rats by magnetic resonance imaging, biochemical imaging, and histology. J Cereb Blood
Flow Metab. 2004;24:789-797.
Hu X, Wester P, Brannstrom T, Watson BD, Gu W. Progressive and reproducible focal cortical ischemia
with or without late spontaneous reperfusion generated by a ring-shaped, laser-driven photothrombotic
lesion in rats. Brain Res Brain Res Protoc. 2001;7:76-85.
Matsuno H, Uematsu T, Umemura K, Takiguchi Y, Asai Y, Muranaka Y, Nakashima M. A simple and
reproducible cerebral thrombosis model in rats induced by a photochemical reaction and the effect of a
plasminogen-plasminogen activator chimera in this model. J Pharmacol Toxicol Methods. 1993;29:165173.
Chen F, Suzuki Y, Nagai N, Peeters R, Sun X, Coudyzer W, Marchal G, Ni Y. Rat cerebral ischemia
induced with photochemical occlusion of proximal middle cerebral artery: a stroke model for MR imaging
research. Magma. 2004;17:103-108.
Chen F, Suzuki Y, Nagai N, Sun X, Wang H, Yu J, Marchal G, Ni Y. Microplasmin and tissue
plasminogen activator: comparison of therapeutic effects in rat stroke model at multiparametric MR
imaging. Radiology. 2007;244:429-438.
Yanagisawa M, Kurihara H, Kimura S, Tomobe Y, Kobayashi M, Mitsui Y, Yazaki Y, Goto K, Masaki T. A
novel potent vasoconstrictor peptide produced by vascular endothelial cells. Nature. 1988;332:411-415.
Robinson MJ, Macrae IM, Todd M, Reid JL, McCulloch J. Reduction of local cerebral blood flow to
pathological levels by endothelin-1 applied to the middle cerebral artery in the rat. Neurosci Lett.
1990;118:269-272.
Sharkey J, Ritchie IM, Kelly PA. Perivascular microapplication of endothelin-1: a new model of focal
cerebral ischaemia in the rat. J Cereb Blood Flow Metab. 1993;13:865-871.
Macrae IM, Robinson MJ, Graham DI, Reid JL, McCulloch J. Endothelin-1-induced reductions in cerebral
blood flow: dose dependency, time course, and neuropathological consequences. J Cereb Blood Flow
Metab. 1993;13:276-284.
Fuxe K, Bjelke B, Andbjer B, Grahn H, Rimondini R, Agnati LF. Endothelin-1 induced lesions of the
frontoparietal cortex of the rat. A possible model of focal cortical ischemia. Neuroreport. 1997;8:26232629.
Nikolova S, Moyanova S, Hughes S, Bellyou-Camilleri M, Lee TY, Bartha R. Endothelin-1 induced
MCAO: dose dependency of cerebral blood flow. J Neurosci Methods. 2009;179:22-28.
Lekic T, Zhang JH. Posterior circulation stroke and animal models. Front Biosci. 2008;13:1827-1844.
Henninger N, Eberius KH, Sicard KM, Kollmar R, Sommer C, Schwab S, Schabitz WR. A new model of
thromboembolic stroke in the posterior circulation of the rat. J Neurosci Methods. 2006;156:1-9.
Pan G, Wright KC. Clot embolic stroke in the vertebrobasilar system of rabbits: a transfemoral
angiographic technique. Cardiovasc Intervent Radiol. 1987;10:285-290.
Gidday JM. Cerebral preconditioning and ischaemic tolerance. Nat Rev Neurosci. 2006;7:437-448.
Koerner IP, Alkayed NJ. Ischemic preconditioning. In: Bhardwaj A, Alkayed NJ, Kirsch JR, Traystman
RJ, eds. Acute stroke, bench to bedside. New York: Informa Helthcare; 2006:345-353.
Perez-Pinzon MA, Xu GP, Dietrich WD, Rosenthal M, Sick TJ. Rapid preconditioning protects rats
against ischemic neuronal damage after 3 but not 7 days of reperfusion following global cerebral
ischemia. J Cereb Blood Flow Metab. 1997;17:175-182.
Nakamura M, Nakakimura K, Matsumoto M, Sakabe T. Rapid tolerance to focal cerebral ischemia in rats
is attenuated by adenosine A1 receptor antagonist. J Cereb Blood Flow Metab. 2002;22:161-170.
Stagliano NE, Perez-Pinzon MA, Moskowitz MA, Huang PL. Focal ischemic preconditioning induces
rapid tolerance to middle cerebral artery occlusion in mice. J Cereb Blood Flow Metab. 1999;19:757-761.
Atochin DN, Clark J, Demchenko IT, Moskowitz MA, Huang PL. Rapid cerebral ischemic preconditioning
in mice deficient in endothelial and neuronal nitric oxide synthases. Stroke. 2003;34:1299-1303.
Kapinya KJ, Lowl D, Futterer C, Maurer M, Waschke KF, Isaev NK, Dirnagl U. Tolerance against
ischemic neuronal injury can be induced by volatile anesthetics and is inducible NO synthase dependent.
Stroke. 2002;33:1889-1898.
Durukan A, Tatlisumak T. Preconditioning-induced ischemic tolerance: a window into endogenous
gearing for cerebroprotection. Exp Transl Stroke Med. 2010;2:2.
Stowe AM, Wacker BK, Cravens PD, Perfater JL, Li MK, Hu R, Freie AB, Stuve O, Gidday JM. CCL2
upregulation triggers hypoxic preconditioning-induced protection from stroke. J Neuroinflammation.
2012;9:33.
Wacker BK, Freie AB, Perfater JL, Gidday JM. Junctional protein regulation by sphingosine kinase 2
contributes to blood-brain barrier protection in hypoxic preconditioning-induced cerebral ischemic
tolerance. J Cereb Blood Flow Metab. 2012;32:1014-1023.
83
253.
254.
255.
256.
257.
258.
259.
260.
261.
262.
263.
264.
265.
266.
267.
268.
269.
270.
271.
272.
273.
274.
275.
276.
277.
278.
279.
280.
Gidday JM, Fitzgibbons JC, Shah AR, Park TS. Neuroprotection from ischemic brain injury by hypoxic
preconditioning in the neonatal rat. Neurosci Lett. 1994;168:221-224.
Miller BA, Perez RS, Shah AR, Gonzales ER, Park TS, Gidday JM. Cerebral protection by hypoxic
preconditioning in a murine model of focal ischemia-reperfusion. Neuroreport. 2001;12:1663-1669.
Bernaudin M, Nedelec AS, Divoux D, MacKenzie ET, Petit E, Schumann-Bard P. Normobaric hypoxia
induces tolerance to focal permanent cerebral ischemia in association with an increased expression of
hypoxia-inducible factor-1 and its target genes, erythropoietin and VEGF, in the adult mouse brain. J
Cereb Blood Flow Metab. 2002;22:393-403.
Lin AM, Dung SW, Chen CF, Chen WH, Ho LT. Hypoxic preconditioning prevents cortical infarction by
transient focal ischemia-reperfusion. Ann N Y Acad Sci. 2003;993:168-178.
Stowe AM, Altay T, Freie AB, Gidday JM. Repetitive hypoxia extends endogenous neurovascular
protection for stroke. Ann Neurol. 2011;69:975-985.
Sharp FR, Ran R, Lu A, Tang Y, Strauss KI, Glass T, Ardizzone T, Bernaudin M. Hypoxic
preconditioning protects against ischemic brain injury. NeuroRx. 2004;1:26-35.
Tang Y, Pacary E, Freret T, Divoux D, Petit E, Schumann-Bard P, Bernaudin M. Effect of hypoxic
preconditioning on brain genomic response before and following ischemia in the adult mouse:
identification of potential neuroprotective candidates for stroke. Neurobiol Dis. 2006;21:18-28.
Kapinya KJ. Ischemic tolerance in the brain. Acta Physiol Hung. 2005;92:67-92.
Baranova O, Miranda LF, Pichiule P, Dragatsis I, Johnson RS, Chavez JC. Neuron-specific inactivation
of the hypoxia inducible factor 1 alpha increases brain injury in a mouse model of transient focal cerebral
ischemia. J Neurosci. 2007;27:6320-6332.
Wacker BK, Park TS, Gidday JM. Hypoxic preconditioning-induced cerebral ischemic tolerance: role of
microvascular sphingosine kinase 2. Stroke. 2009;40:3342-3348.
DeBow SB, Clark DL, MacLellan CL, Colbourne F. Incomplete assessment of experimental
cytoprotectants in rodent ischemia studies. Can J Neurol Sci. 2003;30:368-374.
Du C, Hu R, Csernansky CA, Hsu CY, Choi DW. Very delayed infarction after mild focal cerebral
ischemia: a role for apoptosis? J Cereb Blood Flow Metab. 1996;16:195-201.
Valtysson J, Hillered L, Andine P, Hagberg H, Persson L. Neuropathological endpoints in experimental
stroke pharmacotherapy: the importance of both early and late evaluation. Acta Neurochir (Wien).
1994;129:58-63.
Corbett D, Crooks P. Ischemic preconditioning: a long term survival study using behavioural and
histological endpoints. Brain Res. 1997;760:129-136.
Green AR. Why do neuroprotective drugs that are so promising in animals fail in the clinic? An industry
perspective. Clin Exp Pharmacol Physiol. 2002;29:1030-1034.
Yamaguchi T, Suzuki M, Yamamoto M. YM796, a novel muscarinic agonist, improves the impairment of
learning behavior in a rat model of chronic focal cerebral ischemia. Brain Res. 1995;669:107-114.
Kawamata T, Alexis NE, Dietrich WD, Finklestein SP. Intracisternal basic fibroblast growth factor (bFGF)
enhances behavioral recovery following focal cerebral infarction in the rat. J Cereb Blood Flow Metab.
1996;16:542-547.
Rousselet E, Kriz J, Seidah NG. Mouse model of intraluminal MCAO: cerebral infarct evaluation by
cresyl violet staining. J Vis Exp. 2012
Kloss CU, Thomassen N, Fesl G, Martens KH, Yousri TA, Hamann GF. Tissue-saving infarct volumetry
using histochemistry validated by MRI in rat focal ischemia. Neurol Res. 2002;24:713-718.
Swanson RA, Morton MT, Tsao-Wu G, Savalos RA, Davidson C, Sharp FR. A semiautomated method
for measuring brain infarct volume. J Cereb Blood Flow Metab. 1990;10:290-293.
Goldlust EJ, Paczynski RP, He YY, Hsu CY, Goldberg MP. Automated measurement of infarct size with
scanned images of triphenyltetrazolium chloride-stained rat brains. Stroke. 1996;27:1657-1662.
Tatlisumak T, Takano K, Carano RA, Miller LP, Foster AC, Fisher M. Delayed treatment with an
adenosine kinase inhibitor, GP683, attenuates infarct size in rats with temporary middle cerebral artery
occlusion. Stroke. 1998;29:1952-1958.
Mack WJ, Komotar RJ, Mocco J, Coon AL, Hoh DJ, King RG, Ducruet AF, Ransom ER, Oppermann M,
DeLaPaz R, et al. Serial magnetic resonance imaging in experimental primate stroke: validation of MRI
for pre-clinical cerebroprotective trials. Neurol Res. 2003;25:846-852.
Menzies SA, Hoff JT, Betz AL. Middle cerebral artery occlusion in rats: a neurological and pathological
evaluation of a reproducible model. Neurosurgery. 1992;31:100-106.
Hunter AJ, Hatcher J, Virley D, Nelson P, Irving E, Hadingham SJ, Parsons AA. Functional assessments
in mice and rats after focal stroke. Neuropharmacology. 2000;39:806-816.
D'Hooge R, De Deyn PP. Applications of the Morris water maze in the study of learning and memory.
Brain Res Brain Res Rev. 2001;36:60-90.
Bingham D, Martin SJ, Macrae IM, Carswell HV. Watermaze performance after middle cerebral artery
occlusion in the rat: the role of sensorimotor versus memory impairments. J Cereb Blood Flow Metab.
2012;32:989-999.
Li Y, Chopp M, Chen J, Wang L, Gautam SC, Xu YX, Zhang Z. Intrastriatal transplantation of bone
84
281.
282.
283.
284.
285.
286.
287.
288.
289.
290.
291.
292.
293.
294.
295.
296.
297.
298.
299.
300.
301.
302.
303.
304.
305.
306.
307.
marrow nonhematopoietic cells improves functional recovery after stroke in adult mice. J Cereb Blood
Flow Metab. 2000;20:1311-1319.
Schaar KL, Brenneman MM, Savitz SI. Functional assessments in the rodent stroke model. Exp Transl
Stroke Med. 2010;2:13.
Vannucci SJ, Willing LB, Goto S, Alkayed NJ, Brucklacher RM, Wood TL, Towfighi J, Hurn PD, Simpson
IA. Experimental stroke in the female diabetic, db/db, mouse. J Cereb Blood Flow Metab. 2001;21:52-60.
Sieber FE, Hurn P, Alkayed NJ, Traystman RJ. Gender-based differences in Na+ -K+ adenosine
triphosphatase activity occur in the microcirculation of the diabetic rat brain. Anesthesiology.
2001;94:372-375.
Toung TK, Hurn PD, Traystman RJ, Sieber FE. Estrogen decreases infarct size after temporary focal
ischemia in a genetic model of type 1 diabetes mellitus. Stroke. 2000;31:2701-2706.
Alkayed NJ, Harukuni I, Kimes AS, London ED, Traystman RJ, Hurn PD. Gender-linked brain injury in
experimental stroke. Stroke. 1998;29:159-165.
Shi J, Bui JD, Yang SH, He Z, Lucas TH, Buckley DL, Blackband SJ, King MA, Day AL, Simpkins JW.
Estrogens decrease reperfusion-associated cortical ischemic damage: an MRI analysis in a transient
focal ischemia model. Stroke. 2001;32:987-992.
Leon RL, Li X, Huber JD, Rosen CL. Worsened outcome from middle cerebral artery occlusion in aged
rats receiving 17beta-estradiol. Endocrinology. 2012;153:3386-3393.
Rosen CL, Dinapoli VA, Nagamine T, Crocco T. Influence of age on stroke outcome following transient
focal ischemia. J Neurosurg. 2005;103:687-694.
DiNapoli VA, Huber JD, Houser K, Li X, Rosen CL. Early disruptions of the blood-brain barrier may
contribute to exacerbated neuronal damage and prolonged functional recovery following stroke in aged
rats. Neurobiol Aging. 2008;29:753-764.
Zhang L, Zhang RL, Wang Y, Zhang C, Zhang ZG, Meng H, Chopp M. Functional recovery in aged and
young rats after embolic stroke: treatment with a phosphodiesterase type 5 inhibitor. Stroke.
2005;36:847-852.
Kaur J, Tuor UI, Zhao Z, Barber PA. Quantitative MRI reveals the elderly ischemic brain is susceptible to
increased early blood-brain barrier permeability following tissue plasminogen activator related to claudin
5 and occludin disassembly. J Cereb Blood Flow Metab. 2011;31:1874-1885.
Won SJ, Xie L, Kim SH, Tang H, Wang Y, Mao X, Banwait S, Jin K. Influence of age on the response to
fibroblast growth factor-2 treatment in a rat model of stroke. Brain Res. 2006;1123:237-244.
Kelly KA, Li X, Tan Z, VanGilder RL, Rosen CL, Huber JD. NOX2 inhibition with apocynin worsens
stroke outcome in aged rats. Brain Res. 2009;1292:165-172.
Bardutzky J, Shen Q, Henninger N, Bouley J, Duong TQ, Fisher M. Differences in ischemic lesion
evolution in different rat strains using diffusion and perfusion imaging. Stroke. 2005;36:2000-2005.
Dittmar MS, Vatankhah B, Fehm NP, Schuierer G, Bogdahn U, Horn M, Schlachetzki F. Fischer-344 rats
are unsuitable for the MCAO filament model due to their cerebrovascular anatomy. J Neurosci Methods.
2006;156:50-54.
Duverger D, MacKenzie ET. The quantification of cerebral infarction following focal ischemia in the rat:
influence of strain, arterial pressure, blood glucose concentration, and age. J Cereb Blood Flow Metab.
1988;8:449-461.
Oliff HS, Weber E, Miyazaki B, Marek P. Infarct volume varies with rat strain and vendor in focal cerebral
ischemia induced by transcranial middle cerebral artery occlusion. Brain Res. 1995;699:329-331.
Oliff HS, Weber E, Eilon G, Marek P. The role of strain/vendor differences on the outcome of focal
ischemia induced by intraluminal middle cerebral artery occlusion in the rat. Brain Res. 1995;675:20-26.
Walberer M, Stolz E, Muller C, Friedrich C, Rottger C, Blaes F, Kaps M, Fisher M, Bachmann G, Gerriets
T. Experimental stroke: ischaemic lesion volume and oedema formation differ among rat strains (a
comparison between Wistar and Sprague-Dawley rats using MRI). Lab Anim. 2006;40:1-8.
Sauter A, Rudin M. Strain-dependent drug effects in rat middle cerebral artery occlusion model of stroke.
J Pharmacol Exp Ther. 1995;274:1008-1013.
Oliff HS, Marek P, Miyazaki B, Weber E. The neuroprotective efficacy of MK-801 in focal cerebral
ischemia varies with rat strain and vendor. Brain Res. 1996;731:208-212.
Coyle P. Different susceptibilities to cerebral infarction in spontaneously hypertensive (SHR) and
normotensive Sprague-Dawley rats. Stroke. 1986;17:520-525.
Bailey EL, Smith C, Sudlow CL, Wardlaw JM. Is the spontaneously hypertensive stroke prone rat a
pertinent model of sub cortical ischemic stroke? A systematic review. Int J Stroke. 2011;6:434-444.
Bailey EL, McCulloch J, Sudlow C, Wardlaw JM. Potential animal models of lacunar stroke: a systematic
review. Stroke. 2009;40:e451-458.
Nabika T, Cui Z, Masuda J. The stroke-prone spontaneously hypertensive rat: how good is it as a model
for cerebrovascular diseases? Cell Mol Neurobiol. 2004;24:639-646.
Rees DA, Alcolado JC. Animal models of diabetes mellitus. Diabet Med. 2005;22:359-370.
Lerman LO, Chade AR, Sica V, Napoli C. Animal models of hypertension: an overview. J Lab Clin Med.
2005;146:160-173.
85
308.
309.
310.
311.
312.
313.
314.
315.
316.
317.
318.
319.
320.
321.
322.
323.
324.
325.
326.
327.
328.
329.
330.
331.
332.
333.
334.
335.
336.
Fan X, Qiu J, Yu Z, Dai H, Singhal AB, Lo EH, Wang X. A rat model of studying tissue-type plasminogen
activator thrombolysis in ischemic stroke with diabetes. Stroke. 2012;43:567-570.
Ning R, Chopp M, Yan T, Zacharek A, Zhang C, Roberts C, Cui X, Lu M, Chen J. Tissue plasminogen
activator treatment of stroke in type-1 diabetes rats. Neuroscience. 2012;222:326-332.
Zhu M, Bi X, Jia Q, Shangguan S. The possible mechanism for impaired angiogenesis after transient
focal ischemia in type 2 diabetic GK rats: different expressions of angiostatin and vascular endothelial
growth factor. Biomed Pharmacother. 2010;64:208-213.
MacDougall NJ, Muir KW. Hyperglycaemia and infarct size in animal models of middle cerebral artery
occlusion: systematic review and meta-analysis. J Cereb Blood Flow Metab. 2011;31:807-818.
Huang NC, Wei J, Quast MJ. A comparison of the early development of ischemic brain damage in
normoglycemic and hyperglycemic rats using magnetic resonance imaging. Exp Brain Res.
1996;109:33-42.
Nedergaard M. Transient focal ischemia in hyperglycemic rats is associated with increased cerebral
infarction. Brain Res. 1987;408:79-85.
Liu L, Wang Z, Wang X, Song L, Chen H, Bemeur C, Ste-Marie L, Montgomery J. Comparison of two rat
models of cerebral ischemia under hyperglycemic conditions. Microsurgery. 2007;27:258-262.
Xing Y, Jiang X, Yang Y, Xi G. Hemorrhagic transformation induced by acute hyperglycemia in a rat
model of transient focal ischemia. Acta Neurochir Suppl. 2011;111:49-54.
Ginsberg MD, Busto R. Combating hyperthermia in acute stroke: a significant clinical concern. Stroke.
1998;29:529-534.
van der Worp HB, Sena ES, Donnan GA, Howells DW, Macleod MR. Hypothermia in animal models of
acute ischaemic stroke: a systematic review and meta-analysis. Brain. 2007;130:3063-3074.
Kirsch JR, Traystman RJ, Hurn PD. Anesthetics and cerebroprotection: experimental aspects. Int
Anesthesiol Clin. 1996;34:73-93.
Clarkson AN. Anesthetic-mediated protection/preconditioning during cerebral ischemia. Life Sci.
2007;80:1157-1175.
Ehrlich P. Das Sauerstoffbeduerfnis des organismus: Eine farbenanalytiche Studie. In: Editor, ed.^eds.
Berlin: Hirschwald; 1885.
Hawkins BT, Davis TP. The blood-brain barrier/neurovascular unit in health and disease. Pharmacol
Rev. 2005;57:173-185.
Reese TS, Karnovsky MJ. Fine structural localization of a blood-brain barrier to exogenous peroxidase. J
Cell Biol. 1967;34:207-217.
National Institute of Neurological Disorders and Stroke. Report of the stroke progress review group 2002.
2002
Lo EH, Dalkara T, Moskowitz MA. Mechanisms, challenges and opportunities in stroke. Nat Rev
Neurosci. 2003;4:399-415.
Wardlaw JM, Doubal F, Armitage P, Chappell F, Carpenter T, Munoz Maniega S, Farrall A, Sudlow C,
Dennis M, Dhillon B. Lacunar stroke is associated with diffuse blood-brain barrier dysfunction. Ann
Neurol. 2009;65:194-202.
Oldendorf WH, Cornford ME, Brown WJ. The large apparent work capability of the blood-brain barrier: a
study of the mitochondrial content of capillary endothelial cells in brain and other tissues of the rat. Ann
Neurol. 1977;1:409-417.
Cao Y, Brown SL, Knight RA, Fenstermacher JD, Ewing JR. Effect of intravascular-to-extravascular
water exchange on the determination of blood-to-tissue transfer constant by magnetic resonance
imaging. Magn Reson Med. 2005;53:282-293.
Simionescu M, Simionescu N, Palade GE. Morphometric data on the endothelium of blood capillaries. J
Cell Biol. 1974;60:128-152.
Fenstermacher J, Gross P, Sposito N, Acuff V, Pettersen S, Gruber K. Structural and functional
variations in capillary systems within the brain. Ann N Y Acad Sci. 1988;529:21-30.
Broadwell RD, Salcman M. Expanding the definition of the blood-brain barrier to protein. Proc Natl Acad
Sci U S A. 1981;78:7820-7824.
Butt AM, Jones HC, Abbott NJ. Electrical resistance across the blood-brain barrier in anaesthetized rats:
a developmental study. J Physiol. 1990;429:47-62.
Engelhardt B. Development of the blood-brain barrier. Cell Tissue Res. 2003;314:119-129.
Risau W, Esser S, Engelhardt B. Differentiation of blood-brain barrier endothelial cells. Pathol Biol
(Paris). 1998;46:171-175.
Bodor N, Buchwald P. Targeting of neuropharmaceuticals by chemical delivery systems. In: Dermietzel
R, Spray DC, Nedergaard M, eds. Blood-brain barriers from ontogeny to artificial interfaces. Weinheim:
Wiley-VCH; 2006:463-500.
Dalkara T, Gursoy-Ozdemir Y, Yemisci M. Brain microvascular pericytes in health and disease. Acta
Neuropathol. 2011;122:1-9.
Kamouchi M, Ago T, Kuroda J, Kitazono T. The possible roles of brain pericytes in brain ischemia and
stroke. Cell Mol Neurobiol. 2012;32:159-165.
86
337.
338.
339.
340.
341.
342.
343.
344.
345.
346.
347.
348.
349.
350.
351.
352.
353.
354.
355.
356.
357.
358.
359.
360.
361.
362.
363.
Lindahl P, Johansson BR, Leveen P, Betsholtz C. Pericyte loss and microaneurysm formation in PDGFB-deficient mice. Science. 1997;277:242-245.
Daneman R, Zhou L, Kebede AA, Barres BA. Pericytes are required for blood-brain barrier integrity
during embryogenesis. Nature. 2010;468:562-566.
Yemisci M, Gursoy-Ozdemir Y, Vural A, Can A, Topalkara K, Dalkara T. Pericyte contraction induced by
oxidative-nitrative stress impairs capillary reflow despite successful opening of an occluded cerebral
artery. Nat Med. 2009;15:1031-1037.
del Zoppo GJ. Relationship of neurovascular elements to neuron injury during ischemia. Cerebrovasc
Dis. 2009;27 Suppl 1:65-76.
Hamann GF, Okada Y, Fitridge R, del Zoppo GJ. Microvascular basal lamina antigens disappear during
cerebral ischemia and reperfusion. Stroke. 1995;26:2120-2126.
del Zoppo GJ. The neurovascular unit, matrix proteases, and innate inflammation. Ann N Y Acad Sci.
2010;1207:46-49.
Hamann GF, Okada Y, del Zoppo GJ. Hemorrhagic transformation and microvascular integrity during
focal cerebral ischemia/reperfusion. J Cereb Blood Flow Metab. 1996;16:1373-1378.
Rosenberg GA, Estrada EY, Dencoff JE. Matrix metalloproteinases and TIMPs are associated with
blood-brain barrier opening after reperfusion in rat brain. Stroke. 1998;29:2189-2195.
Gasche Y, Fujimura M, Morita-Fujimura Y, Copin JC, Kawase M, Massengale J, Chan PH. Early
appearance of activated matrix metalloproteinase-9 after focal cerebral ischemia in mice: a possible role
in blood-brain barrier dysfunction. J Cereb Blood Flow Metab. 1999;19:1020-1028.
Asahi M, Wang X, Mori T, Sumii T, Jung JC, Moskowitz MA, Fini ME, Lo EH. Effects of matrix
metalloproteinase-9 gene knockout on the proteolysis of blood-brain barrier and white matter
components after cerebral ischemia. J Neurosci. 2001;21:7724-7732.
Fukuda S, Fini CA, Mabuchi T, Koziol JA, Eggleston LL, Jr., del Zoppo GJ. Focal cerebral ischemia
induces active proteases that degrade microvascular matrix. Stroke. 2004;35:998-1004.
Sood RR, Taheri S, Candelario-Jalil E, Estrada EY, Rosenberg GA. Early beneficial effect of matrix
metalloproteinase inhibition on blood-brain barrier permeability as measured by magnetic resonance
imaging countered by impaired long-term recovery after stroke in rat brain. J Cereb Blood Flow Metab.
2008;28:431-438.
Yang Y, Estrada EY, Thompson JF, Liu W, Rosenberg GA. Matrix metalloproteinase-mediated disruption
of tight junction proteins in cerebral vessels is reversed by synthetic matrix metalloproteinase inhibitor in
focal ischemia in rat. J Cereb Blood Flow Metab. 2007;27:697-709.
Liu J, Jin X, Liu KJ, Liu W. Matrix metalloproteinase-2-mediated occludin degradation and caveolin-1mediated claudin-5 redistribution contribute to blood-brain barrier damage in early ischemic stroke stage.
J Neurosci. 2012;32:3044-3057.
Gu Y, Zheng G, Xu M, Li Y, Chen X, Zhu W, Tong Y, Chung SK, Liu KJ, Shen J. Caveolin-1 regulates
nitric oxide-mediated matrix metalloproteinases activity and blood-brain barrier permeability in focal
cerebral ischemia and reperfusion injury. J Neurochem. 2012;120:147-156.
Ramos-Fernandez M, Bellolio MF, Stead LG. Matrix metalloproteinase-9 as a marker for acute ischemic
stroke: a systematic review. J Stroke Cerebrovasc Dis. 2011;20:47-54.
Kniesel U, Wolburg H. Tight junctions of the blood-brain barrier. Cell Mol Neurobiol. 2000;20:57-76.
Bauer HC, Traweger A, Bauer H. Proteins of the tight junction in the blood-brain barrier. In: Sharma HS,
Westman J, eds. Blood-spinal cord and brain barriers in health and disease. San Diego, California:
Elsevier Academic Press; 2004:1-10.
Wolburg H. The endothelial frontier. In: Dermietzel R, Spray DC, Nedergaard M, eds. Blood-brain
barriers from ontogeny to artificial interfaces. Weinheim: Wiley-VCH; 2006:77-107.
Rosenberg GA, Yang Y. Vasogenic edema due to tight junction disruption by matrix metalloproteinases
in cerebral ischemia. Neurosurg Focus. 2007;22:E4.
Sandoval KE, Witt KA. Blood-brain barrier tight junction permeability and ischemic stroke. Neurobiol Dis.
2008;32:200-219.
Furuse M, Hirase T, Itoh M, Nagafuchi A, Yonemura S, Tsukita S, Tsukita S. Occludin: a novel integral
membrane protein localizing at tight junctions. J Cell Biol. 1993;123:1777-1788.
Saitou M, Fujimoto K, Doi Y, Itoh M, Fujimoto T, Furuse M, Takano H, Noda T, Tsukita S. Occludindeficient embryonic stem cells can differentiate into polarized epithelial cells bearing tight junctions. J
Cell Biol. 1998;141:397-408.
Balda MS, Whitney JA, Flores C, Gonzalez S, Cereijido M, Matter K. Functional dissociation of
paracellular permeability and transepithelial electrical resistance and disruption of the apical-basolateral
intramembrane diffusion barrier by expression of a mutant tight junction membrane protein. J Cell Biol.
1996;134:1031-1049.
Kago T, Takagi N, Date I, Takenaga Y, Takagi K, Takeo S. Cerebral ischemia enhances tyrosine
phosphorylation of occludin in brain capillaries. Biochem Biophys Res Commun. 2006;339:1197-1203.
Correale J, Villa A. Cellular elements of the blood-brain barrier. Neurochem Res. 2009;34:2067-2077.
Nitta T, Hata M, Gotoh S, Seo Y, Sasaki H, Hashimoto N, Furuse M, Tsukita S. Size-selective loosening
87
364.
365.
366.
367.
368.
369.
370.
371.
372.
373.
374.
375.
376.
377.
378.
379.
380.
381.
382.
383.
384.
385.
386.
387.
388.
389.
of the blood-brain barrier in claudin-5-deficient mice. J Cell Biol. 2003;161:653-660.
Martin-Padura I, Lostaglio S, Schneemann M, Williams L, Romano M, Fruscella P, Panzeri C,
Stoppacciaro A, Ruco L, Villa A, et al. Junctional adhesion molecule, a novel member of the
immunoglobulin superfamily that distributes at intercellular junctions and modulates monocyte
transmigration. J Cell Biol. 1998;142:117-127.
Stevenson BR, Siliciano JD, Mooseker MS, Goodenough DA. Identification of ZO-1: a high molecular
weight polypeptide associated with the tight junction (zonula occludens) in a variety of epithelia. J Cell
Biol. 1986;103:755-766.
Nico B, Ribatti D. Morphofunctional aspects of the blood-brain barrier. Curr Drug Metab. 2012;13:50-60.
Mishiro K, Ishiguro M, Suzuki Y, Tsuruma K, Shimazawa M, Hara H. A broad-spectrum matrix
metalloproteinase inhibitor prevents hemorrhagic complications induced by tissue plasminogen activator
in mice. Neuroscience. 2012;205:39-48.
Dejana E, Orsenigo F, Lampugnani MG. The role of adherens junctions and VE-cadherin in the control
of vascular permeability. J Cell Sci. 2008;121:2115-2122.
Simard M, Nedergaard M. The neurobiology of glia in the context of water and ion homeostasis.
Neuroscience. 2004;129:877-896.
Zhou J, Kong H, Hua X, Xiao M, Ding J, Hu G. Altered blood-brain barrier integrity in adult aquaporin-4
knockout mice. Neuroreport. 2008;19:1-5.
Manley GT, Fujimura M, Ma T, Noshita N, Filiz F, Bollen AW, Chan P, Verkman AS. Aquaporin-4
deletion in mice reduces brain edema after acute water intoxication and ischemic stroke. Nat Med.
2000;6:159-163.
Zeng XN, Xie LL, Liang R, Sun XL, Fan Y, Hu G. AQP4 knockout aggravates ischemia/reperfusion injury
in mice. CNS Neurosci Ther. 2012;18:388-394.
Crone C. The Permeability Of Capillaries In Various Organs As Determined By Use Of The 'Indicator
Diffusion' Method. Acta Physiol Scand. 1963;58:292-305.
Nag S. Blood-brain barrier permeability using tracers and immunohistochemistry. Methods Mol Med.
2003;89:133-144.
Hoffmann A, Bredno J, Wendland M, Derugin N, Ohara P, Wintermark M. High and Low Molecular
Weight Fluorescein Isothiocyanate (FITC)-Dextrans to Assess Blood-Brain Barrier Disruption: Technical
Considerations. Transl Stroke Res. 2011;2:106-111.
Xu Q, Qaum T, Adamis AP. Sensitive blood-retinal barrier breakdown quantitation using Evans blue.
Invest Ophthalmol Vis Sci. 2001;42:789-794.
Manaenko A, Chen H, Kammer J, Zhang JH, Tang J. Comparison Evans Blue injection routes:
Intravenous versus intraperitoneal, for measurement of blood-brain barrier in a mice hemorrhage model.
J Neurosci Methods. 2011;195:206-210.
Tofts PS, Berkowitz BA. Measurement of capillary permeability from the Gd enhancement curve: a
comparison of bolus and constant infusion injection methods. Magn Reson Imaging. 1994;12:81-91.
Kastrup A, Groschel K, Ringer TM, Redecker C, Cordesmeyer R, Witte OW, Terborg C. Early disruption
of the blood-brain barrier after thrombolytic therapy predicts hemorrhage in patients with acute stroke.
Stroke. 2008;39:2385-2387.
Hjort N, Wu O, Ashkanian M, Solling C, Mouridsen K, Christensen S, Gyldensted C, Andersen G,
Ostergaard L. MRI detection of early blood-brain barrier disruption: parenchymal enhancement predicts
focal hemorrhagic transformation after thrombolysis. Stroke. 2008;39:1025-1028.
Kassner A, Mandell DM, Mikulis DJ. Measuring permeability in acute ischemic stroke. Neuroimaging Clin
N Am. 2011;21:315-325.
Saria A, Lundberg JM. Evans blue fluorescence: quantitative and morphological evaluation of vascular
permeability in animal tissues. J Neurosci Methods. 1983;8:41-49.
Uyama O, Okamura N, Yanase M, Narita M, Kawabata K, Sugita M. Quantitative evaluation of vascular
permeability in the gerbil brain after transient ischemia using Evans blue fluorescence. J Cereb Blood
Flow Metab. 1988;8:282-284.
Hawkins BT, Egleton RD. Fluorescence imaging of blood-brain barrier disruption. J Neurosci Methods.
2006;151:262-267.
Ohno K, Pettigrew KD, Rapoport SI. Lower limits of cerebrovascular permeability to nonelectrolytes in
the conscious rat. Am J Physiol. 1978;235:H299-307.
Huang ZG, Xue D, Preston E, Karbalai H, Buchan AM. Biphasic opening of the blood-brain barrier
following transient focal ischemia: effects of hypothermia. Can J Neurol Sci. 1999;26:298-304.
Masada T, Hua Y, Xi G, Ennis SR, Keep RF. Attenuation of ischemic brain edema and cerebrovascular
injury after ischemic preconditioning in the rat. J Cereb Blood Flow Metab. 2001;21:22-33.
Mayhan WG, Heistad DD. Permeability of blood-brain barrier to various sized molecules. Am J Physiol.
1985;248:H712-718.
Jin AY, Tuor UI, Rushforth D, Kaur J, Muller RN, Petterson JL, Boutry S, Barber PA. Reduced blood
brain barrier breakdown in P-selectin deficient mice following transient ischemic stroke: a future
therapeutic target for treatment of stroke. BMC Neurosci. 2010;11:12.
88
390.
391.
392.
393.
394.
395.
396.
397.
398.
399.
400.
401.
402.
403.
404.
405.
406.
407.
408.
409.
410.
411.
412.
413.
414.
Dimitrijevic OB, Stamatovic SM, Keep RF, Andjelkovic AV. Effects of the chemokine CCL2 on bloodbrain barrier permeability during ischemia-reperfusion injury. J Cereb Blood Flow Metab. 2006;26:797810.
Schmidt-Kastner R, Szymas J, Hossmann KA. Immunohistochemical study of glial reaction and serumprotein extravasation in relation to neuronal damage in rat hippocampus after ischemia. Neuroscience.
1990;38:527-540.
Klohs J, Steinbrink J, Bourayou R, Mueller S, Cordell R, Licha K, Schirner M, Dirnagl U, Lindauer U,
Wunder A. Near-infrared fluorescence imaging with fluorescently labeled albumin: a novel method for
non-invasive optical imaging of blood-brain barrier impairment after focal cerebral ischemia in mice. J
Neurosci Methods. 2009;180:126-132.
Abulrob A, Brunette E, Slinn J, Baumann E, Stanimirovic D. In vivo optical imaging of ischemic bloodbrain barrier disruption. Methods Mol Biol. 2011;763:423-439.
Tofts PS. Modeling tracer kinetics in dynamic Gd-DTPA MR imaging. J Magn Reson Imaging. 1997;7:91101.
Patlak CS, Blasberg RG, Fenstermacher JD. Graphical evaluation of blood-to-brain transfer constants
from multiple-time uptake data. J Cereb Blood Flow Metab. 1983;3:1-7.
Ewing JR, Knight RA, Nagaraja TN, Yee JS, Nagesh V, Whitton PA, Li L, Fenstermacher JD. Patlak
plots of Gd-DTPA MRI data yield blood-brain transfer constants concordant with those of 14C-sucrose in
areas of blood-brain opening. Magn Reson Med. 2003;50:283-292.
Hoffmann A, Bredno J, Wendland MF, Derugin N, Hom J, Schuster T, Su H, Ohara PT, Young WL,
Wintermark M. Validation of in vivo magnetic resonance imaging blood-brain barrier permeability
measurements by comparison with gold standard histology. Stroke. 2011;42:2054-2060.
Kassner A, Roberts TP, Moran B, Silver FL, Mikulis DJ. Recombinant tissue plasminogen activator
increases blood-brain barrier disruption in acute ischemic stroke: an MR imaging permeability study.
AJNR Am J Neuroradiol. 2009;30:1864-1869.
Vidarsson L, Thornhill RE, Liu F, Mikulis DJ, Kassner A. Quantitative permeability magnetic resonance
imaging in acute ischemic stroke: how long do we need to scan? Magn Reson Imaging. 2009;27:12161222.
Kuroiwa T, Ting P, Martinez H, Klatzo I. The biphasic opening of the blood-brain barrier to proteins
following temporary middle cerebral artery occlusion. Acta Neuropathol (Berl). 1985;68:122-129.
Belayev L, Busto R, Zhao W, Ginsberg MD. Quantitative evaluation of blood-brain barrier permeability
following middle cerebral artery occlusion in rats. Brain Res. 1996;739:88-96.
Candelario-Jalil E, Taheri S, Yang Y, Sood R, Grossetete M, Estrada EY, Fiebich BL, Rosenberg GA.
Cyclooxygenase inhibition limits blood-brain barrier disruption following intracerebral injection of tumor
necrosis factor-alpha in the rat. J Pharmacol Exp Ther. 2007;323:488-498.
Rosenberg GA. Neurological diseases in relation to the blood-brain barrier. J Cereb Blood Flow Metab.
2012;32:1139-1151.
Wagner GF, Hampong M, Park CM, Copp DH. Purification, characterization, and bioassay of teleocalcin,
a glycoprotein from salmon corpuscles of Stannius. Gen Comp Endocrinol. 1986;63:481-491.
Chang AC, Janosi J, Hulsbeek M, de Jong D, Jeffrey KJ, Noble JR, Reddel RR. A novel human cDNA
highly homologous to the fish hormone stanniocalcin. Mol Cell Endocrinol. 1995;112:241-247.
Chang AC, Reddel RR. Identification of a second stanniocalcin cDNA in mouse and human:
stanniocalcin 2. Mol Cell Endocrinol. 1998;141:95-99.
Yeung BH, Law AY, Wong CK. Evolution and roles of stanniocalcin. Mol Cell Endocrinol. 2012;349:272280.
Yeung HY, Lai KP, Chan HY, Mak NK, Wagner GF, Wong CK. Hypoxia-inducible factor-1-mediated
activation of stanniocalcin-1 in human cancer cells. Endocrinology. 2005;146:4951-4960.
Liu D, Huang L, Wang Y, Wang W, Wehrens XH, Belousova T, Abdelrahim M, DiMattia G, SheikhHamad D. Human stanniocalcin-1 suppresses angiotensin II-induced superoxide generation in
cardiomyocytes through UCP3-mediated anti-oxidant pathway. PLoS One. 2012;7:e36994.
Sheikh-Hamad D. Mammalian stanniocalcin-1 activates mitochondrial antioxidant pathways: new
paradigms for regulation of macrophages and endothelium. Am J Physiol Renal Physiol. 2010;298:F248254.
Ito D, Walker JR, Thompson CS, Moroz I, Lin W, Veselits ML, Hakim AM, Fienberg AA, Thinakaran G.
Characterization of stanniocalcin 2, a novel target of the mammalian unfolded protein response with
cytoprotective properties. Mol Cell Biol. 2004;24:9456-9469.
Franzen AM, Zhang KZ, Westberg JA, Zhang WM, Arola J, Olsen HS, Andersson LC. Expression of
stanniocalcin in the epithelium of human choroid plexus. Brain Res. 2000;887:440-443.
Ratkovic S, Wagner GF, Ciriello J. Distribution of stanniocalcin binding sites in the lamina terminalis of
the rat. Brain Res. 2008;1218:141-150.
Li K, Dong D, Yao L, Dai D, Gu X, Guo L. Identification of STC1 as an beta-amyloid activated gene in
human brain microvascular endothelial cells using cDNA microarray. Biochem Biophys Res Commun.
2008;376:399-403.
89
415.
416.
417.
418.
419.
420.
421.
422.
423.
424.
425.
426.
427.
428.
429.
430.
431.
432.
433.
434.
435.
436.
437.
438.
439.
Zhang KZ, Westberg JA, Paetau A, von Boguslawsky K, Lindsberg P, Erlander M, Guo H, Su J, Olsen
HS, Andersson LC. High expression of stanniocalcin in differentiated brain neurons. Am J Pathol.
1998;153:439-445.
Zhang K, Lindsberg PJ, Tatlisumak T, Kaste M, Olsen HS, Andersson LC. Stanniocalcin: A molecular
guard of neurons during cerebral ischemia. Proc Natl Acad Sci U S A. 2000;97:3637-3642.
Long Y, Zou L, Liu H, Lu H, Yuan X, Robertson CS, Yang K. Altered expression of randomly selected
genes in mouse hippocampus after traumatic brain injury. J Neurosci Res. 2003;71:710-720.
Westberg JA, Serlachius M, Lankila P, Penkowa M, Hidalgo J, Andersson LC. Hypoxic preconditioning
induces neuroprotective stanniocalcin-1 in brain via IL-6 signaling. Stroke. 2007;38:1025-1030.
Byun JS, Lee JW, Kim SY, Kwon KJ, Sohn JH, Lee K, Oh D, Kim SS, Chun W, Lee HJ. Neuroprotective
effects of stanniocalcin 2 following kainic acid-induced hippocampal degeneration in ICR mice. Peptides.
2010;31:2094-2099.
Holmes DI, Zachary IC. Vascular endothelial growth factor regulates stanniocalcin-1 expression via
neuropilin-1-dependent regulation of KDR and synergism with fibroblast growth factor-2. Cell Signal.
2008;20:569-579.
Chen C, Jamaluddin MS, Yan S, Sheikh-Hamad D, Yao Q. Human stanniocalcin-1 blocks TNF-alphainduced monolayer permeability in human coronary artery endothelial cells. Arterioscler Thromb Vasc
Biol. 2008;28:906-912.
Manalo DJ, Rowan A, Lavoie T, Natarajan L, Kelly BD, Ye SQ, Garcia JG, Semenza GL. Transcriptional
regulation of vascular endothelial cell responses to hypoxia by HIF-1. Blood. 2005;105:659-669.
Chang AC, Cha J, Koentgen F, Reddel RR. The murine stanniocalcin 1 gene is not essential for growth
and development. Mol Cell Biol. 2005;25:10604-10610.
Pedrono E, Durukan A, Strbian D, Marinkovic I, Shekhar S, Pitkonen M, Abo-Ramadan U, Tatlisumak T.
An optimized mouse model for transient ischemic attack. J Neuropathol Exp Neurol. 2010;69:188-195.
Nagaraja TN, Karki K, Ewing JR, Divine GW, Fenstermacher JD, Patlak CS, Knight RA. The MRImeasured arterial input function resulting from a bolus injection of Gd-DTPA in a rat model of stroke
slightly underestimates that of Gd-[14C]DTPA and marginally overestimates the blood-to-brain influx rate
constant determined by Patlak plots. Magn Reson Med. 2010;63:1502-1509.
Tatlisumak T, Carano RA, Takano K, Opgenorth TJ, Sotak CH, Fisher M. A novel endothelin antagonist,
A-127722, attenuates ischemic lesion size in rats with temporary middle cerebral artery occlusion: a
diffusion and perfusion MRI study. Stroke. 1998;29:850-857.
Abramoff MD, Magalhaes PJ, Ram SJ. Image processing with Imaje J. Biophotonics Int. 2004;11:36-42.
Strbian D, Karjalainen-Lindsberg ML, Tatlisumak T, Lindsberg PJ. Cerebral mast cells regulate early
ischemic brain swelling and neutrophil accumulation. J Cereb Blood Flow Metab. 2006;26:605-612.
Patlak CS, Blasberg RG. Graphical evaluation of blood-to-brain transfer constants from multiple-time
uptake data. Generalizations. J Cereb Blood Flow Metab. 1985;5:584-590.
Murray CJ, Lopez AD. Mortality by cause for eight regions of the world: Global Burden of Disease Study.
Lancet. 1997;349:1269-1276.
Seshadri S, Beiser A, Kelly-Hayes M, Kase CS, Au R, Kannel WB, Wolf PA. The lifetime risk of stroke:
estimates from the Framingham Study. Stroke. 2006;37:345-350.
Bose B, Osterholm JL, Berry R. A reproducible experimental model of focal cerebral ischemia in the cat.
Brain Res. 1984;311:385-391.
Kastrup A, Engelhorn T, Beaulieu C, de Crespigny A, Moseley ME. Dynamics of cerebral injury,
perfusion, and blood-brain barrier changes after temporary and permanent middle cerebral artery
occlusion in the rat. J Neurol Sci. 1999;166:91-99.
Neumann-Haefelin T, Kastrup A, de Crespigny A, Yenari MA, Ringer T, Sun GH, Moseley ME. Serial
MRI after transient focal cerebral ischemia in rats: dynamics of tissue injury, blood-brain barrier damage,
and edema formation. Stroke. 2000;31:1965-1972; discussion 1972-1963.
Lennmyr F, Ericsson A, Gerwins P, Ahlstrom H, Terent A. Increased brain injury and vascular leakage
after pretreatment with p38-inhibitor SB203580 in transient ischemia. Acta Neurol Scand. 2003;108:339345.
Nagel S, Wagner S, Koziol J, Kluge B, Heiland S. Volumetric evaluation of the ischemic lesion size with
serial MRI in a transient MCAO model of the rat: comparison of DWI and T1WI. Brain Res Brain Res
Protoc. 2004;12:172-179.
Nagel S, Su Y, Horstmann S, Heiland S, Gardner H, Koziol J, Martinez-Torres FJ, Wagner S.
Minocycline and hypothermia for reperfusion injury after focal cerebral ischemia in the rat: effects on
BBB breakdown and MMP expression in the acute and subacute phase. Brain Res. 2008;1188:198-206.
Veltkamp R, Siebing DA, Sun L, Heiland S, Bieber K, Marti HH, Nagel S, Schwab S, Schwaninger M.
Hyperbaric oxygen reduces blood-brain barrier damage and edema after transient focal cerebral
ischemia. Stroke. 2005;36:1679-1683.
Lin CY, Chang C, Cheung WM, Lin MH, Chen JJ, Hsu CY, Chen JH, Lin TN. Dynamic changes in
vascular permeability, cerebral blood volume, vascular density, and size after transient focal cerebral
ischemia in rats: evaluation with contrast-enhanced magnetic resonance imaging. J Cereb Blood Flow
90
440.
441.
442.
443.
444.
445.
446.
447.
448.
449.
Metab. 2008;28:1491-1501.
Merten CL, Knitelius HO, Assheuer J, Bergmann-Kurz B, Hedde JP, Bewermeyer H. MRI of acute
cerebral infarcts, increased contrast enhancement with continuous infusion of gadolinium.
Neuroradiology. 1999;41:242-248.
del Zoppo GJ, Schmid-Schonbein GW, Mori E, Copeland BR, Chang CM. Polymorphonuclear
leukocytes occlude capillaries following middle cerebral artery occlusion and reperfusion in baboons.
Stroke. 1991;22:1276-1283.
Zhang RL, Chopp M, Chen H, Garcia JH. Temporal profile of ischemic tissue damage, neutrophil
response, and vascular plugging following permanent and transient (2H) middle cerebral artery occlusion
in the rat. J Neurol Sci. 1994;125:3-10.
Wu XD, Du LN, Wu GC, Cao XD. Effects of electroacupuncture on blood-brain barrier after cerebral
ischemia-reperfusion in rat. Acupunct Electrother Res. 2001;26:1-9.
Pillai DR, Dittmar MS, Baldaranov D, Heidemann RM, Henning EC, Schuierer G, Bogdahn U,
Schlachetzki F. Cerebral ischemia-reperfusion injury in rats--a 3 T MRI study on biphasic blood-brain
barrier opening and the dynamics of edema formation. J Cereb Blood Flow Metab. 2009;29:1846-1855.
Lin TN, Sun SW, Cheung WM, Li F, Chang C. Dynamic changes in cerebral blood flow and angiogenesis
after transient focal cerebral ischemia in rats. Evaluation with serial magnetic resonance imaging. Stroke.
2002;33:2985-2991.
Albayrak S, Zhao Q, Siesjo BK, Smith ML. Effect of transient focal ischemia on blood-brain barrier
permeability in the rat: correlation to cell injury. Acta Neuropathol. 1997;94:158-163.
Nagaraja TN, Keenan KA, Brown SL, Fenstermacher JD, Knight RA. Relative distribution of plasma flow
markers and red blood cells across BBB openings in acute cerebral ischemia. Neurol Res. 2007;29:7880.
Nagaraja TN, Karki K, Ewing JR, Croxen RL, Knight RA. Identification of variations in blood-brain barrier
opening after cerebral ischemia by dual contrast-enhanced magnetic resonance imaging and T 1sat
measurements. Stroke. 2008;39:427-432.
Nagaraja TN, Keenan KA, Fenstermacher JD, Knight RA. Acute leakage patterns of fluorescent plasma
flow markers after transient focal cerebral ischemia suggest large openings in blood-brain barrier.
Microcirculation. 2008;15:1-14.
91
ORIGINAL PUBLICATIONS
92

Benzer belgeler

Evaluation of effects of memantine on cerebral ischemia in rats

Evaluation of effects of memantine on cerebral ischemia in rats and mitochondrial dysfunction is believed to be one of the most generalized causes favoring the development of neurodegenerative diseases.18 Memantine was evaluated in animals against primary insul...

Detaylı